Sei sulla pagina 1di 136

Development of Predictive Formulae for the A1 Temperature in Creep Strength Enhanced

Ferritic Steels

THESIS

Presented in Partial Fulfillment of the Requirements for the Degree Master of Science in
the Graduate School of The Ohio State University

By

Lun Wang, B.S.

Graduate Program in Welding Engineering

The Ohio State University

2010

Master's Examination Committee:

Dr. John C. Lippold, Advisor

Dr. Sudarsanam S. Babu


Copyright by

Lun Wang

2010
Abstract

The creep strength enhanced ferritic (CSEF) steels P91 and P92 are extensively

used in fossil supercritical power plants due to their improved creep strength at elevated

temperatures. Loss of creep strength and/or toughness may occur in CSEF steel welds

due to formation of fresh martensite, ferrite, or retained austenite during welding and post

weld heat treatment (PWHT). Predicting the critical phase transformation temperatures is

of practical importance for the development of appropriate welding and PWHT

procedures. The available empirical formula proposed by Andrew [1] for the

determination of the A1 temperature does not cover the composition range of CSEF steels

and the effect of carbon and nitrogen has not been taken into account. Thermodynamic

simulation software such as Thermo-Calc and JMat-Pro could predict equilibrium

transformation temperatures (i.e. equilibrium formation of austenite, ferrite in steels).

However, the microstructure in CSEF steels which is of practical importance for

determination of the A1 temperature is tempered martensite (in as delivered condition).

Santella [65] developed two formulae for P91 and P92 steels based on thermodynamic

simulation data recently. Predictions by Santella’s formulae tend to be conservative

estimates of the A1 temperature in P91 and P92 steels [65].

Thermodynamic simulations and experimental measurement based Design of

Experiment (DOE) approaches and fractional data analysis were applied to develop

formulae for predicting the A1 temperature in P91 and P92 steels. The alloying elements

ii
with a significant effect on the A1 temperature in P91 steels screened out by the

thermodynamic simulation based DOE approach are: Ni, Mn, Si, N, Cr, Mo, C, V. No

interactions were found between these significant alloying elements. A measurement

based DOE was developed based on the elements with significant effect on the A1

temperature determined by the thermodynamic simulation based DOE. The test samples

were melted using a button melting system in argon environment. These samples were

subjected to homogenizing, rolling, normalizing, and tempering in order to reproduce to

the manufacturing process of commercial CSEF steels. The A1 temperatures in these

samples were measured by Single Sensor Differential Thermal Analysis (SS-DTA). The

results of the measurement based DOE were processed using fractional data analysis to

develop predictive formulae for the A1 temperature in P91 and P92 steels. It was found

that C, N, Si and Cr could influence the A1 temperature largely; Ni and Mn were not the

only determining factor of the A1 temperature. The relations between the concentrations

of most alloying elements and the A1 temperature of P91 and P92 steels were quadratic

function. The predictions of these formulae were validated by comparison to measured

values of the A1 temperature in P91 and P92 steels by dilatometry.

iii
Dedication

This document is dedicated to my family.

iv
Acknowledgments

A special thanks is extended to Dr. Boian Alexandrov and Dr. John Lippold for

their guidance over the last two years. It has been a great experience and I have learned

so much from both of you. Thanks to Dr. Sudarsanam Babu for your inspirations of deep

thought.

I would like to acknowledge Prof. Theodore Allen of ISE department for his help

with software JMP8 and data analysis.

Thanks to all the members of Welding and Joining Metallurgy Group, I will miss

our lunch time conversations.

This project was funded by American Electric Power and Babcock & Wilcox. I

am grateful for their support. Additional chemical analysis was sponsored by Babcock &

Wilcox. Lastly, I want to thank John Siefert for his valuable support and suggestions

throughout this time.

v
Vita

March 1988 ...................................................Born – Anqing, China

2004 ...............................................................High School Diploma, Susong High School,

Anqing, Anhui Province, China

2008 ...............................................................B.S. Welding Engineering, Harbin Institute

of Technology, China

2008 to present ..............................................Graduate Research Associate, Welding

Engineering Program, The Ohio State

University

Publications

Lun Wang, Boian Alexandrov and John Lippold, “Predicting A1 Temperature in CSEF

Steels”, MS&T’10, ASM International, Houston, TX, October 2010

Fields of Study

Major Field: Welding Engineering

vi
Table of Contents

Abstract ............................................................................................................................... ii

Dedication .......................................................................................................................... iv

Acknowledgments............................................................................................................... v

Vita..................................................................................................................................... vi

Publications ........................................................................................................................ vi

Fields of Study ................................................................................................................... vi

Table of Contents .............................................................................................................. vii

List of Tables ..................................................................................................................... xi

List of Figures .................................................................................................................. xiv

Chapter 1 INTRODUCTION .............................................................................................. 1

Chapter 2 LITERATURE REVIEW ................................................................................... 4

2.1 Creep Strength Enhanced Ferritic (CSEF) Steels ..................................................... 4

2.1.1 History of CSEF Steels .......................................................................................... 4

2.1.1.1 P91 Steel .......................................................................................................... 6

2.1.1.2 P92 Steel .......................................................................................................... 8

vii
2.1.1.3 P122 Steel ........................................................................................................ 9

2.1.2 Microstructure of CSEF Steels ............................................................................. 11

2.1.2.1 Basic Phases in CSEF Steels ......................................................................... 11

2.1.2.2 Precipitates..................................................................................................... 13

2.1.3 Effect of Alloying Elements ................................................................................. 17

2.1.3.1 Carbon ........................................................................................................... 17

2.1.3.2 Chromium ...................................................................................................... 20

2.1.3.3 Nickel............................................................................................................. 21

2.1.3.4 Boron ............................................................................................................. 22

2.1.3.5 Nitrogen ......................................................................................................... 24

2.1.3.6 Molybdenum .................................................................................................. 26

2.1.3.7 Tungsten ........................................................................................................ 27

2.1.3.8 Molybdenum versus Tungsten....................................................................... 29

2.1.3.9 Copper ........................................................................................................... 31

2.1.3.10 Vanadium and Niobium............................................................................... 32

2.1.3.11 Cobalt........................................................................................................... 35

2.1.3.12 Tantalum ...................................................................................................... 35

2.1.3.13 Manganese and Silicon ................................................................................ 36

2.2 Phase Transformation Analysis Techniques ........................................................... 37

viii
2.2.1 Differential Thermal Analysis (DTA) .................................................................. 37

2.2.2 Differential Scanning Calorimetry (DSC)............................................................ 38

2.2.3 Dilatometry........................................................................................................... 39

2.2.4 Single Sensor Differential Thermal Analysis (SS-DTA) ..................................... 41

2.2.5 The Electrothermomechanical Tester (ETMT) .................................................... 43

2.3 Predicting Methods for Critical Temperatures ........................................................ 44

2.3.1 Predictive Formulae for Ms and Mf Temperatures ............................................... 44

2.3.2 Predictive Formulae for Bs and Bf Temperatures ................................................. 49

2.3.3 Predictive Formulae for Ac1 and Ac3 Temperatures ............................................ 51

Chapter 3 OBJECTIVES .................................................................................................. 52

Chapter 4 MATERIALS AND EXPERIMENTAL PROCEDURES............................... 53

4.1 Design of Experiments ....................................................................................... 53

4.1.1 Model Based DOE ............................................................................................... 53

4.1.2 Measurement based DOE and separate experimental study ................................ 62

4.2 Sample Preparation ............................................................................................ 63

4.2.1 Materials ............................................................................................................... 63

4.2.2 Sample Preparation .............................................................................................. 64

4.3 Evaluation of Composition Control ................................................................... 67

4.4 Determination of the A1 temperature ................................................................. 69

ix
4.5 Validation of formula predicted A1 temperature ................................................ 71

Chapter 5 RESULTS AND DISCUSSION ...................................................................... 72

5.1 Model Based DOE ............................................................................................. 72

5.2 Experimental Results.......................................................................................... 76

5.2.1 Measurement based DoE for P91 Steel ................................................................ 76

5.2.2 Determination of the effect of carbon and sillicon on the A1 temperature in P91

Steel ............................................................................................................................... 87

5.2.3 Determination of the effect of nitrogen on the A1 temperature in P91 Steel ....... 90

5.2.3 Determination of the effect of molybdenum and tungsten on the A1 temperature

in P92 Steel.................................................................................................................... 91

5.2.4 Determination of the effect of copper on the A1 temperature in P122 Steel ........ 94

5.3 Measurement based DOE formulae for the A1 Temperatures in CSEF Steels .. 96

5.3.1 Predictive formula for the A1 temperature in P91 Steel ....................................... 96

5.3.2 Predictive formula for the A1 temperature in P92 Steel ..................................... 100

5.4 Validation of formula predicted A1 temperature .............................................. 105

Chapter 6 CONCLUSIONS ............................................................................................ 108

References ....................................................................................................................... 111

x
List of Tables

Table 2.1: Composition of specified 9Cr Steels (wt.%) [11] [12]. ..................................... 7

Table 2.2: Development of Ms temperature predictive formulae [61].............................. 46

Table 2.3: Applicable composition range of formulae for Ms (wt.%) [1][56]~[63]. ........ 48

Table 2.4: Formulae for Bs and Bf [60] [62] [64].............................................................. 50

Table 2.5: Applicable composition range of formulae for Bs and Bf (wt.%) [60] [62] [64].

........................................................................................................................................... 50

Table 2.6: Formulae for Ac1 and Ac3 [1]. ......................................................................... 51

Table 2.7: Applicable composition range of formulae for Ac1 and Ac3 (wt.%) [1]. ........ 51

Table 4.1: ASTM A335 specification for P91, P92 and P122 steels (wt.%). ................... 58

Table 4.2: The first part of model based DOE table with JMat-Pro predicted A1

temperature values (wt.%). ............................................................................................... 59

Table 4.3: The second part of model based DOE table with JMat-Pro predicted A1

temperature values (wt.%). ............................................................................................... 60

Table 4.4: Purity of pure elements. ................................................................................... 63

Table 4.5: Composition of base materials and 1080 steel wire (wt.%). ........................... 63

Table 4.6: Compositions of melted and heat treated samples (wt.%). .............................. 68

Table 4.7: Composition and SS-DTA measured A1 temperature of P91 steel specimens

(wt.%)................................................................................................................................ 71
xi
Table 5.1: Composition range of measurement based DoE for P91 steels (wt.%). .......... 79

Table 5.2: Chemical analysis of measurement based DOE samples (wt.%). ................... 85

Table 5.3: Application range of Equation (2) (wt.%). ...................................................... 86

Table 5.4: Measured composition of P91 steel samples with varied C concentration

(wt.%)................................................................................................................................ 89

Table 5.5: Measured composition of P91 steel samples with varied Si concentration

(wt.%)................................................................................................................................ 89

Table 5.6: Composition of commercial and laboratory heats of P91 steel samples used for

determination of the effect of nitrogen on the A1 temperature (wt.%). ............................ 90

Table 5.7: Measured composition of P92 steel samples with varied W concentration

(wt.%)................................................................................................................................ 93

Table 5.8: Measured composition of P92 steel samples with varied Mo concentration... 93

Table 5.9: Measured composition of P122 steel samples with varied Cu concentration. 95

Table 5.10: Composition (wt.%) of 10 commercial P91 steels and corresponding

measured (by dilatometry) and formula predicted A1 temperature. ................................. 98

Table 5.11: Predictions for P91 steel samples from different available models (°C). ...... 99

Table 5.12: Composition (wt.%) of 19 commercial P92 steels and corresponding

measured (by dilatometry) and formula predicted A1 temperature. ............................... 103

Table 5.13: Predictions for P92 steel samples from different available models (°C). .... 104

Table 5.14: Composition and hardness of tempered P91 specimens. ............................. 106

xii
xiii
List of Figures

Figure 1.1: Deviations of models predicted A1 temperature from measured A1 temperature

............................................................................................................................................. 3

Figure 2.1: Development history of 9-12Cr steels [8] ........................................................ 5

Figure 2.2: Applied stress to rupture time curves of NF616 and P91 [12]. ........................ 8

Figure 2.3: Influence of alloying elements on properties in HCM12A [5] ...................... 10

Figure 2.4: Typical microstructures of P91 steel: a) fully tempered martensite; b)

tempered martensite and δ-ferrite [4]................................................................................ 11

Figure 2.5: Influence of δ-ferrite on toughness and creep strength: a) Relation between

percentage of δ-ferrite and toughness at room temperature; b) Relation between

percentage of δ-ferrite and creep-rupture time [5] [6]. ..................................................... 12

Figure 2.6: Typical precipitates and their distribution in CSEF steels [3] [22]. ............... 13

Figure 2.7: Growth of precipitate particles in CSEF steels [3]. ........................................ 14

Figure 2.8: Growth of Laves phase particles in P92 steel at (a) 600°C; (b) 650°C [3]. ... 15

Figure 2.9: Time to rupture and minimum creep rate of 9Cr steel at 600°C and 140 Mpa

as a function of carbon concentration [18] ....................................................................... 18

Figure 2.10: Amount of M23C6 carbide and MX carbonitrides in 9Cr steel at tempering

temperature of 800°C as a function of carbon concentration [18] .................................... 18

xiv
Figure 2.11: Y-groove restraint weld cracking test of 11Cr-2W-0.4Mo-1Cu-v-Nb-0.003B

steels with different carbon concentrations at different preheating temperature [5] ........ 19

Figure 2.12: Influence of Ni and Cu on the extrapolated creep rupture strength of 0.1C-

11Cr-2W-0.4Mo-Cu steels [5] .......................................................................................... 21

Figure 2.13: Formation process of M23(CB)6 during heat treatment [18]. ........................ 23

Figure 2.14: (a) Enrichment of boron in M23C6 near PAGBs in 9Cr-3W-3Co-0.2V-0.08C

steel with 139 ppm boron; (b) Size of M23C6 particles as a function of boron

concentration [18]. ............................................................................................................ 23

Figure 2.15: Time to rupture and minimum creep rate of 9Cr-3W-3Co-0.2V-0.05Nb-

0.08C steel with 140ppm B at 650°C and 120Mpa as a function of nitrogen concentration

[18]. ................................................................................................................................... 25

Figure 2.16: Creep rupture strength of 9Cr-0.5Mo-W steels tested at 600°C and 700°C

[28]. ................................................................................................................................... 28

Figure 2.17: Tensile properties of 9Cr-0.5Mo-W steels tested at room temperature [28].28

Figure 2.19: Effect of W+Mo on creep rupture strength and Charpy absorbed energy in

12Cr turbine steels [32]..................................................................................................... 30

Figure 2.18: (a) Effect of Mo+W on creep rupture strength of 12% Cr steels [27]; (b)

Effect of Mo and W on 105h creep rupture strength at 650°C and Charpy absorbed energy

at 20°C of 12% Cr turbine steels (0.13C-0.05Si-0.5Mn-11.2Cr-0.8Ni-0.2V-0.5Nb-0.05N-

Mo-W) [31]. ...................................................................................................................... 30

Figure 2.20: (a) Effect of V and Nb on 104h creep rupture strength of NF616 at 600°C;

(2) Effect of V and Nb on 104h creep rupture strength of NF616 at 650°C [30]. ............ 32

xv
Figure 2.21: (a) Influence of V and Nb on tensile properties ultimate tensile strength at

823K; (b) Influence of V and Nb on ductile-brittle transition temperature after aging at

873K for 6000h [34]. ........................................................................................................ 33

Figure 2.22: Influence of Nb on creep rupture time at 873K [34]. ................................... 34

Figure 2.23: Influence of V on creep rupture time at 873K [34]. ..................................... 34

Figure 2.24: (a) Schematic of DTA installation [48]; (b) Typical DTA curves of low

alloy Cr-Mo steel a- hot rolled and b- as quenched [49]. ................................................. 37

Figure 2.25: (a) Schematic of DSC installation; (b) Typical DSC result. [50]................. 38

Figure 2.26: Linear thermal expansion coefficient of ferrite and austenite [51]. ............. 39

Figure 2.27: Typical Dilatometry curve of steel [ASTM A 1033-04]. ............................. 40

Figure 2.28: Schematic of SS-DTA installation [54]........................................................ 42

Figure 2.29: SS-DTA result of steel HSLA-65: (a) sample cooling curve and reference

curve; (b) difference between sample cooling curve and reference curve as a function of

sample temperature [54]. .................................................................................................. 42

Figure 2.30: Image of ETMT [55] .................................................................................... 43

Figure 2.31: Formula of Ms developed by Beres and Irmer(1994) [63]. .......................... 47

Figure 4.1: Structure of the model based DOE used in this study. ................................... 54

Figure 4.2: JMP8 starter menu and DOE features. ........................................................... 55

Figure 4.3: Defining the response for model based DOE in JMP8. ................................. 55

Figure 4.4: Defining the factors for model based DOE in JMP8...................................... 56

Figure 4.5: Defining the model effects for model based DOE in JMP8. .......................... 57

xvi
Figure 4.6: Defining the number of center points and replicates for model based DOE in

JMP8. ................................................................................................................................ 58

Figure 4.7: The model fitting platform for model based DOE in JMP8. .......................... 61

Figure 4.8: Schematics of sample melting, casting and heat treatment: a) Tungsten arc

button melting system; b) Sample casting system; c) Cubic copper casting mold; d) Shape

of sample after casting; e) Shape of sample after hot-rolling; f); Heat history of

normalizing; g) Heat history of tempering....................................................................... 66

Figure 4.9: EDX map analysis result of heat treated samples. ......................................... 68

Figure 4.10: A2, A1, A3 temperatures in laboratory melted heat of P91 steel determined by

SS-DTA............................................................................................................................. 70

Figure 5.1: The JMat-Pro predicted A1 temperature versus A1 temperature predicted by

the model based DOE. ...................................................................................................... 73

Figure 5.2: Elements with significant influence on the JMat-Pro predicted A1 temperature

in P91 steel and their estimated effects. ............................................................................ 74

Figure 5.3: Prediction profiler of model based DOE model of A1 temperature in P91 steel.

........................................................................................................................................... 74

Figure 5.4: Interaction profiler of model based DOE model of A1 temperature in P91

steel. .................................................................................................................................. 75

Figure 5.5: Defining the response for measurement based DOE in JMP8. ...................... 77

Figure 5.6: Defining the factors for measurement based DOE in JMP8. ......................... 78

Figure 5.7: Defining the model for measurement based DOE in JMP8. .......................... 78

xvii
Figure 5.8: Defining the number of center points and replicates for measurement based

DOE in JMP8. ................................................................................................................... 79

Figure 5.9: Measurement based DOE table with measured A1 temperature values (testing

error of type K thermocouple is ± 0.4%). ......................................................................... 80

Figure 5.10: Predictive model of the A1 temperature in P91 steel as a function of

composition of Ni, Mn, Cr, Mo, V and Cu developed using JMP8 by measurement based

DOE study......................................................................................................................... 82

Figure 5.12: Elements with significant influence on measured A1 temperature in P91 steel

and their estimate effects. ................................................................................................. 83

Figure 5.11: The actual (measured) A1 temperature versus the experiment based DOE

model predicted A1 temperature ........................................................................................ 83

Figure 5.13: Prediction profiler of measurement based DOE model of A1 temperature in

P91 steel. ........................................................................................................................... 84

Figure 5.14: Influence of C and Si on the A1 temperature in P91 steel. ........................... 88

Figure 5.15: Influence of W and Mo on the A1 temperature in P92 steel......................... 92

Figure 5.16: Influence of Cu on the A1 temperature in P122 steel. .................................. 95

Figure 5.17: Comparison of A1 temperature in P91 steels predicted by measurement

based DOE formula and A1 temperature determined by dilatometry. ............................... 97

Figure 5.18: Predictions for P91 steel samples from different available models. ............ 99

Figure 5.19: Comparison of A1 temperature in P92 steels predicted by the proposed

measurement based DOE formula and A1 temperature determined by dilatometry. ...... 101

Figure 5.20: Predictions for P92 steel samples from different available models. .......... 104

xviii
Figure 5.21: Microscopy of tempered samples (X1000): a) Specimen1 tempered at 825°C

for 30min, hardness is 140 HV; b) Specimen1 tempered at 850°C for 30min, hardness is

297 HV; c) Micro-hardness of specimen1 tempered at 850°C (load 50g); d) Specimen2

tempered at 825°C for 30min, hardness is 152 HV; e) Specimen2 tempered at 850°C for

30min, hardness is 238 HV; f) Micro-hardness of specimen2 tempered at 850°C (load

50g). ................................................................................................................................ 107

xix
Chapter 1 INTRODUCTION

In order to improve the thermal efficiency of fossil power plants and reduce the

emission of carbon dioxide, CSEF steels (P91 and P92) have been used extensively. In

these steels, (Cr, Fe, Mo)23C6 and fine niobium and vanadium carbonitrides (MX) form

within a martensite matrix during the tempering treatment. Because of the stability at

elevated temperature, the precipitation of vanadium nitride plays a decisive role in

maintaining creep strength at high temperature [2]. Degradation occurs during the creep

process, due to coarsening of precipitates and growth of Laves phase (Fe,Cr)2Mo [2].

Compared to P91 steel, fine stable Laves phase particles (Fe,Cr)2(Mo,W) were found in

P92 steel due to the addition of tungsten [3]. A small amount of boron (of about 0.001

wt.%) has a strong stabilizing effect on M23C6 particles [3]. To guarantee the high

temperature creep strength, the microstructure of CSEF steels should be tempered

martensite with properly distributed carbide and carbonitride precipitates [4]. Tempering

above the A1 temperature results in formation of ferrite and fresh martensite, and

dissolution /coarsening of the carbides, which adversely affects toughness and creep

strength of CSEF steels [5] [6].

Accurate determination of the A1 temperature in CSEF steels is essential for the

development of proper tempering procedures. The A1 temperature in CSEF steels is

strongly affected by the steel composition and there is no data available. Andrew [1]

1
proposed a formula for the Ac1 temperature as shown below, based on collected data of

196 general steels with a wide composition range:

Ac1(˚C) = 723 - 10.7Mn - 16.9Ni + 29.1Si + 16.9Cr + 290As + 6.38W

This function does not consider the influence of C, N, Mo, V, and Cu on the A1

temperature, and its application range does not cover the content of Cr, Cu in CSEF steels.

The formula proposed by Andrew shows high deviation from A1 temperatures in CSEF

steels measured by dilatometry and SS DTA. Thermodynamic simulation software such

as Thermo-Calc and JMat-Pro also could be used to predict equilibrium phase

transformation temperatures in steels. However, by comparing to our data of measured A1

temperatures, A1 temperature calculated by thermodynamics calculation with JMat-Pro

software was found to be lower than measured A1 temperature in CSEF steels. Based

linear regression analysis of thermodynamic simulation data base, Santella [65]

developed two expressions which were supposed to predict the A1 temperature in P91 and

P92 steels thermodynamically:

Grade 91 steel: A1(˚C) = 805 + 2.5Cr + 18.1Mo + 19.1Si + 37.1V + 19.2Nb – 67.3C
– 130.6N – 60.5Mn – 72.3Ni

Grade 92 steel: A1(˚C) = 778 + 4.9Cr + 22.6Mo +10.8W + 22.9Si + 43.6V + 20.2Nb
– 80.6C – 150.7N – 55.1Mn – 68.0Ni

2
The deviations of different models predicted A1 temperature from measured A1

temperature (heating rate is 28˚C/h) in 8 commercial P91 steels are shown in Figure 1.1.

65
55
45
35 Andrew’s Function
25 JMat-Pro
15 Thermo-Calc
5 Michael's Expression
-5
ZC131 ZC132 ZC ZC134 ANP8 SRP1 ZC133 ZC11
-15
-25

Figure 1.1: Deviations of models predicted A1 temperature from measured A1 temperature

This paper describes the development of formulae for predicting the A1

temperature in CSEF steels. These formulae have been developed using a DOE approach

based on SS-DTA measurements of the A1 temperature in laboratory heats of P91 and

P92 steels. The proposed formulae have been validated by comparison to measured

values of the A1 temperature in P91 and P92 steels. These formulae show better

agreement with the A1 temperature in commercial CSEF steel measured by SS DTA and

dilatometry than the available predictive formulae and models.


3
Chapter 2 LITERATURE REVIEW

2.1 Creep Strength Enhanced Ferritic (CSEF) Steels

2.1.1 History of CSEF Steels

According to Scarlin[7], power plant steels should meet the following general

requirements: high 100,000h creep-rupture strength (about 100Mpa) at elevated

temperatures of 600°C or even higher; good weldability for pipes and tubes; fabricable

for large components; highly resistant to high temperature steam oxidation and corrosion

for boiler tubes; uniform hardening and mechanical properties in large rotor components;

high toughness and low susceptibility to embrittlement and softening when service at

elevated temperatures for long time.

The development history of 9-12Cr ferritic steels is shown in Figure 2.1 [8]; the

creep-rupture strength at 600°C and 105 hours is believed to be the most important

mechanical property for power plant steels. In the 1940s [9], 2.25Cr-1Mo steel were first

used as a high chromium steel in power plants. Until the 1950s [10], 2.25Cr-1Mo steel

was the most used high chromium steel in Europe and the US; however the poor creep

resistance of 2.25Cr-1Mo steel was an issue for some of components. In order to achieve

higher thermal efficiency of fossil-fuel power plants and reduce the CO2 emission, new 9-

12Cr ferritic steels with higher creep-rupture strength at elevated temperatures and

improved oxidation and corrosion resistance were developed in Europe, Japan and the US

[10] [11] [12].


4
(P92)

(P122)

Figure 2.1: Development history of 9-12Cr steels [8]

5
2.1.1.1 P91 Steel

In the middle 1960s [11], the first 9Cr steel was developed by Center for

Metallurgical Research of Liege in Belgium. Several years later in France, the first

material named EM 12 was fabricated and was approved for power plant applications. As

shown in Table 2.1, EM 12 contains 9 wt.% chromium and 2 wt.% molybdenum. Since

1978, EM 12 had been installed in lots of fossil fuel power boilers mainly in Europe and

some other countries and many are still in use today. High ferrite stabilizers in EM12

result in a duplex microstructure containing δ-ferrite and low impact toughness. So EM12

is not applicable for thick components. In the 1970s [10], the Oak Ridge National

Laboratory (ORNL) in the US paid much attention to developing a modified 9Cr steel.

The composition of modified 9Cr steel (T91) is shown in Table 2.1. Composition of EM

12 was optimized by ORNL, modified 9Cr-1Mo steel contains lower manganese,

molybdenum, niobium and vanadium, and nitrogen of controlled extent was added for the

first time. In the early 1980s, ASTM approved the modified 9Cr-1Mo steel for power

plant applications in standard A 213 T91 and A 335 P91. Excellent corrosion resistance

of T91 and P91 steels meets the requirement of modern materials to be applied in modern

fossil-fuel power boilers with operating temperature of about 600°C to 625°C.

6
Grades EM12 T91 ASTM NF616 P92 ASTM
A 213 A 335
C 0.15 max. 0.08-0.12 0.07-0.10 0.07-0.13

Mn 0.08-1.30 0.30-0.60 0.42-0.47 0.30-0.60

Si 0.20-0.65 0.20-0.50 0.06-0.10 0.50 max.

S Max. 0.030 0.010 Same as 0.010


9Cr-2Mo steel
P Max. 0.030 0.020 Same as 0.020
9Cr-2Mo steel
Cr 8.50-10.50 8.00-9.50 8.98-9.05 8.50-9.50

Ni 0.3 max. 0.40 max. Same as 0.40 max.


9Cr-2Mo steel
Mo 1.70-2.30 0.85-1.05 0.49-0.52 0.30-0.60

V 0.20-0.40 0.18-0.25 0.16-0.21 0.15-0.25

Nb 0.30-0.55 0.06-0.10 0.049-0.060 0.04-0.09

N - 0.030-0.070 0.037-0.048 0.030-0.070

W - - 1.65-1.72 1.50-2.00

B - - - 10-60ppm

Al - 0.040 max. - 0.040 max.

Sn 0.030 max. - - -

Table 2.1: Composition of specified 9Cr Steels (wt.%) [11] [12].

7
2.1.1.2 P92 Steel

In 1993 [12], Nippon Steel developed a 9Cr-0.5Mo-1.8W-Nb-V steel, named

NF616. Composition range of first four experimentally tested NF616 steels is shown in

Table 2.1. The composition of NF616 samples were the same as T91 steel except that

part of molybdenum was replaced by tungsten in order to obtain solid-solution

strengthening. Vanadium and niobium were added to ensure precipitation strengthening

with carbonitrides. In NF616 microstructure of fully tempered martensite without δ-

ferrite was achieved. As shown in Figure 2.2, at 600°C the slope of applied stress to

rupture time curve of NF616 was gentler than P91 steel, and the creep-rupture strength at

600°C and 105 hours of NF616 was significantly higher than P91 steel. In the 1990s, the

NF616 was standardized by ASTM as pipe steel (P92) and as tube steel (T92) [13] [14].

Figure 2.2: Applied stress to rupture time curves of NF616 and P91 [12].
8
2.1.1.3 P122 Steel

In the late 1960s [10], in order to solve the problem with duplex phase EM12, the

first 12Cr-1Mo steel was developed in Germany with the designation of X20CrMoV12-1

(X20). While X20 had a fully martensitic microstructure, it exhibits lower creep strength

than EM12 at elevated temperatures and was difficult to weld, primarily due to high

carbon content of 0.20 wt.% [15]. In the 1980’s [16], a 0.1C-12Cr-1Mo-1W-V-Nb steel

tube (HCM12) was manufactured by Sumitomo Metal in Japan, the creep-rupture

strength of both base metal and welded joint of HCM12 was higher than P91 and T91

steels. However, for thick components, formation of δ-ferrite in HCM12 decreases

impact toughness [5]. The influence of each alloying element on the properties of

HCM12 steel is shown in Figure 2.3. In 1992 [5], by optimizing the composition of

HCM12, a 0.1C-11Cr-2W-0.4Mo-1Cu steel pipe (HCM12A) was produced by Sumitomo

Metal in Japan. Copper was added to prohibit the formation of δ-ferrite, and also make it

less susceptible to the HAZ softening. Adding more tungsten instead of molybdenum,

and around 0.002 wt.% boron, effectively improved the creep property of that steel. In

1994 [14], HCM12A was standardized by ASME Code as P122.

9
Figure 2.3: Influence of alloying elements on properties in HCM12A [5]

10
2.1.2 Microstructure of CSEF Steels

2.1.2.1 Basic Phases in CSEF Steels

In order to guarantee the creep strength at high temperature (600°C to 650°C), the

optimal microstructure of CSEF steels is tempered martensite [4]. Figure 2.4(a) shows the

optimal microstructure of fully temepred martensite for P91 steel; Figure 2.4(b) shows

the existence of δ-ferrite in P91 steel. Testing results showed sample (a) had both better

tensile strength and creep-rupture strength at 650°C than sample (b).

a) b)

Figure 2.4: Typical microstructures of P91 steel: a) fully tempered martensite; b)


tempered martensite and δ-ferrite [4].

It was found that the presence of δ-ferrite in the microstructure degraded both

toughness and creep strength of CSEF steels [5] [6]. As shown in Figure 2.5(a), the

11
toughness of a 0.1C-11Cr-W-Mo-V steel decreased with the content of δ-ferrite

increasing: Figure 2.5(b) showed the increasing of δ-ferrite formation reduced the rupture

time of a 0.1C-9Cr-1Mo-V steel at 600°C and 167Mpa.

b)
a)

Figure 2.5: Influence of δ-ferrite on toughness and creep strength: a) Relation between
percentage of δ-ferrite and toughness at room temperature; b) Relation between
percentage of δ-ferrite and creep-rupture time [5] [6].

The reason of δ-ferrite degrading creep strength of CSEF steels was researched by

Kimura et al. [45]. It was found that significant C and N concentration gradients between

tempered martensite and δ-ferrite stimulated formation of some large precipitate particles

thus degraded the creep strength of the steel.

12
2.1.2.2 Precipitates

The main advantage of CSEF steels is their high creep strength at elevated

temperature (600°C to 650°C), which requires that the steel could keep high dislocation

density and fine subgrain microstructure during the long-term creep process at 600°C to

650°C [3]. It was found that precipitation hardening was the main hardening mechanism

in CSEF steels; fine stable precipitations particles could pin the dislocations and subgrain

boundaries. As shown in Figure 2.6, the main precipitates in CSEF steels are M23C6

carbides, MX carbonitrides and Laves phase. The M23C6 carbides and MX carbonitrides

could form during tempering process, Laves phase only forms during the high

temperature creep service. The M23C6 carbides and Laves phase primarily form at prior

austenite boundaries and lath boundaries, while fine MX carbonitrides precipitate in

ferrite matrix within the laths [22].

Figure 2.6: Typical precipitates and their distribution in CSEF steels [3] [22].

13
a)

P91
Laves phase
P92

P91
M23C6
P92

b)

Figure 2.7: Growth of precipitate particles in CSEF steels [3].

J. Hald collected data and developed diagram of precipitate particles growing in

P91 and P92 steels [2]. As shown in Figure 2.7(a), at 600°C the growth of M23C6 carbides

in P92 steel was much slower than that of P91 steel, this was because of addition of boron

in P92 steel. Figure 2.7(b) shows that the MX carbonitride particles are stable during

aging at 600°C up to 60,000 hours. Figure 2.7(a) also shows that after aging at 600°C for

50,000 hours the mean radius of Laves phase particles in P91 is five times larger than that

of P91 steel. This was believed to be the main reason for P92 steel having higher creep

strength than P91 steel. Since the aging temperature (600°C) is close to the dissolution

temperature of Mo-rich Laves phase (650°C) in P91 steel, the nucleation of Mo-rich
14
Laves phase became very difficult and only a few particles nucleated and then grew into

extended large particles. The W-rich Laves phase in P92 steel had a higher dissolution

temperature (720°C) than Mo-rich Laves phase. This explanation also was verified by the

growth of Laves phase particles in P92 steel at 650°C, as shown in Figure 2.8. At 650°C

the Laves phase in P92 steel could grow into much bigger particles at relatively shorter

time than at 600°C, because the aging temperature of 650°C is too close to the dissolution

temperature of W-rich Laves phase (720°C) [3].

Figure 2.8: Growth of Laves phase particles in P92 steel at (a) 600°C; (b) 650°C [3].

A study on 11Cr-0.87Cu steel (P122) showed that copper precipitate particles

formed during tempering process grew fast during aging process [27]. It was also

observed that Laves phase particles in P122 steel were finer than these in P92 steel. This

was explained by the presence of more nucleation sites for Laves phase in P122 steel

because the copper precipitates could serve as nucleation sites for Laves phase [27].
15
In 1986 Z-phase (Cr(V, Nb)N) was first observed by Schnabel et al. in steel

X19CrMoVNbN11-1 which contained 12 wt.% chromium [46]. The niobium content of

this 12Cr steel was around five times higher than CSEF steels, so Z-phase was not

expected to form in CSEF steels. However, recently Z-phase in form of Cr(V, Nb)N was

found in a T122 steel with 12 wt.% chromium after exposure at elevated temperature for

more than 10,000 hours, and this caused breakdown of long-term stable creep strength of

this CSEF steel [3]. It was explained by the formation of Z-phase which consumed V, Nb

and N from the MX carbonitrides and thus dissolved fine MX carbonitrides particles that

were essential for long-term creep strength of CSEF steels. It was believed that the high

chromium content accelerated the formation of Z-phase. This could be a potential

problem for niobium containing steel with more than 10 wt.% chromium [3]. This

became a challenge for the development of new CSEF steels with higher oxidation

resistance that required higher chromium concentration. Z-phase in form of Cr(V, Ta)N

was also observed in a steel containing 12 wt. % chromium and tantalum.

In 2009, Hald and Danielsen proposed the idea of Z-phase strengthened 12Cr

steel [47]. Accelerating the precipitation of Z-phase particles and forming fine Z-phase

could make the Z-phase particles induce precipitating strengthening to steel. This could

be achieved by increasing the chromium to as high as possible (12 wt.% for CSEF steels)

and adding cobalt. It was also calculated that Z-phase in form of CrNbN and CrTaN

grows at much slower rate than CrVN or Cr(V, Nb)N, even at the same level as the MN

nitrides. Onging work in TU Denmark showed that it is possible to developed high

chromium Z-phase (main population is below 30nm) strengthened matensitic steels [47].

16
2.1.3 Effect of Alloying Elements

2.1.3.1 Carbon

Copeland and Licina’s research revealed that increasing the carbon content from

0.01 wt.% to 0.13 wt.% increased the creep and rupture strength in 2.25Cr-1Mo steel.

They also found that as the temperature increased, the effect of carbon on creep and

rupture strength decreased [17]. Masuyama et al. proposed that in modified 9Cr weld

metal a minimum of 0.08 wt.% carbon content is required to guarantee the creep strength

[9]. The creep strength in steel with low carbon content is reduced because of the

formation of insufficient amount of M23C6 and MC6 precipitates.

However, as shown in Figure 2.9, recent research by Fujio Abe showed that the

9Cr-3W-3Co-0.2V-0.05Nb-0.05N steel with lower than 0.05 wt.% carbon had higher

time to rupture and lower minimum creep rate at 600°C and 140Mpa [18]. The rupture

time was independent of carbon concentration when carbon concentration was higher

than 0.047 wt.%, but it significantly increased when carbon concentrations was below

0.018 wt.%. This agreed with the Thermo-calc evaluation of equilibrium phases as shown

in Figure 2.10, the creep strength of 9Cr steel was significantly improved by elimination

of M23C6 and dispersion of nanosize MX nitrides. Wey et al. reported that the coarsening

rate of Cr carbides in iron was much higher than carbonitrides of V and Nb at elevated

temperatures [19]. The amount of M23C6 carbides rich in Cr decreased with decreasing

carbon concentration. There were more MX carbonitrides than M23C6 carbides at carbon

concentrations below 0.02 wt.% during tempering at 800 °C. The MX carbonitrides

consisted of mainly vanadium nitrides and of small amount of niobium nitrides [20].

17
Figure 2.9: Time to rupture and minimum creep rate of 9Cr steel at 600°C and 140 Mpa
as a function of carbon concentration [18]

Figure 2.10: Amount of M23C6 carbide and MX carbonitrides in 9Cr steel at tempering
temperature of 800°C as a function of carbon concentration [18]

18
Carbon concentration also had significant effect on weldability of 9-12Cr steels.

Iseda et al. carried on y-groove restraint weld cracking test for 11Cr-2W-0.4Mo-1Cu-v-

Nb-0.003B steels with different carbon concentrations at different preheating

temperatures, the result is shown in Figure 2.11 [5]. The result of y-groove restraint weld

cracking test showed the susceptibility to cold weld cracking decreased by lowering the

carbon concentration from 0.08 wt.% to 0.14 wt.%. The cracking was completely

suppressed when the preheating temperature was adjusted above 150°C and carbon

concentration was controlled less than 0.12 wt.%. Increasing of the carbon concentration

could also result in decreasing the toughness [10].

Figure 2.11: Y-groove restraint weld cracking test of 11Cr-2W-0.4Mo-1Cu-v-Nb-0.003B


steels with different carbon concentrations at different preheating temperature [5]

19
2.1.3.2 Chromium

Chromium is one of the major alloying elements in Cr-Mo ferritic steels to

promote oxidation and corrosion resistance [10]. With addition of chromium, a stable

layer of oxide of stoichiometry (Fe,Cr)2O3 forms on the steel surface, which protects steel

from further oxidation and corrosion [21]. Experience shows that 12 wt.% chromium are

needed to raise the oxidation resistance to an acceptable level [42]. Recent research by

Abellan et al. showed that the oxidation resistance of their samples with 10.8 wt.%

chromium and 11 wt.% chromium was distinctively better than of piping steel P92 in an

environment of Ar-50%H2O at 600°C and 650°C, and even somewhat higher than sample

with 12 wt.% chromium [43].

Although chromium does not exhibit an obvious effect on creep strength, high

strength is more likely to be obtained when chromium is 2 wt.% and between 9 and 12

wt.% in a ferritic steel, strength declines between these two chromium concentrations

[22]. However, the tested creep rupture time at 650°C and 100Mpa of newly developed

ferritic-martensitic steels during last decades proved that highest creep strength was

obtained at about 9.5 wt.% chromium [42]. Chromium is a strong carbide former; the

most common Cr-rich carbide are M23C6 carbides; “M” is predominantly Cr but also have

Fe, Mo and W. The M23C6 carbides are beneficial to the rupture strength when they are

finely dispersed, however they are no stable at elevated temperature because they

spheroidize easily and become large and blocky quickly [10]. Chromium is also a strong

ferrite stabilizer, when adding chromium into Cr-Mo steel there is a risk of formation of

δ-ferrite.

20
2.1.3.3 Nickel

Nickel is a strong austenite stabilizer that could inhibit the formation of δ-ferrite

to improve the toughness of steels. Nickel also could decrease the A1 temperature [22].

The effect of nickel on creep rupture strength at 600°C for 105 hours is shown in Figure

2.12 [5]. Nickel decreased the creep rupture strength at 600°C for 105 hours quickly when

its concentration is above 0.4 wt.%; it was believed that Laves phase precipitation was

accelerated by excessive nickel addition. The ASTM specification limits the nickel

content in both P91 and P92 steels bellow 0.4 wt.%.

Figure 2.12: Influence of Ni and Cu on the extrapolated creep rupture strength of 0.1C-
11Cr-2W-0.4Mo-Cu steels [5]

21
2.1.3.4 Boron

Boron has a strong stabilizing effect on M23C6 carbides. Takahashi et al. reported

that for 0.2C-10.5Cr-1.5Mo-0.2V-0.2Nb-0.02N steel, addiction of 120 to 370ppm boron

improved the creep rupture strength at 650°C [23]. They observed that the M23C6 carbides

at the prior austenite grain boundaries (PAGBs) were finer than in a steel without boron.

The influence of boron on M23C6 carbides in a 9Cr-3W-3Co-0.2V-0.08C steel (without

N) was recently researched by Fujio Abe [18]. B is a strong nitride former, so nitrogen

was not added to avoid formation of boron nitride. Therefore only M23C6 carbides and no

MX cabonitrides formed after tempering. As shown in Figure 2.13, during the

normalizing process grain boundary segregation of boron took place. During the

tempering process precipitation of M23C6 carbide could take place preferentially at grain

boundaries. The M23C6 enriched with B become M23(CB)6 at the vicinity of PAGBs. As

shown in Figure 2.14(a), after heat treatment and aging the segregation of B could be up

to around 5 at.%. Figure 2.14(b) shows that after ageing at 625°C for 10300 hours, the

size of M23C6 carbides on the PAGBs is reduced at higher B concentration. Reduction of

coarsening rate of M23C6 carbides on PAGBs was considered to be related to the

segregation of boron, but the detailed mechanism was not clarified. In Europe, 9-12Cr

steel containing 100ppm or higher B were considered to be used in next-generation USC

power plants at 625°C to 650°C [24] [25].

22
Figure 2.13: Formation process of M23(CB)6 during heat treatment [18].

(ppm)

Figure 2.14: (a) Enrichment of boron in M23C6 near PAGBs in 9Cr-3W-3Co-0.2V-0.08C


steel with 139 ppm boron; (b) Size of M23C6 particles as a function of boron concentration
[18].

23
2.1.3.5 Nitrogen

Nitrogen is more effective than carbon in improving creep rupture strength in

ferritic steels [6]. Nitrogen has low solubility in the ferrite and tends to form precipitate

as fine MX carbonitrides. However, other large nitrides particles like Cr2N and BN

should be avoided to prevent deterioration of toughness [18] [26]. As shown in Figure

2.15, for a 9Cr-3W-3Co-0.2V-0.05Nb-0.08C steel with 140ppm boron the peak time to

rupture and minimum creep rate at 650°C and 120Mpa were located at around 80 to 100

ppm nitrogen [18]. By calculation, at normalizing temperature of about 1100°C in 9Cr-

3W-3Co-0.2V-0.05Nb-0.08C steel with 140ppm boron, only around 95 ppm nitrogen

could dissolve in the matrix without formation of boron nitride. Addition of small amount

of nitrogen without formation of boron nitride during normalizing significantly improved

the creep strength of steel. However, formation of boron nitride when more nitrogen was

added caused the degradation of creep strength of steel. It was also reported by Sawada et

al. that large Cr2N particles could form along the lath, block and pocket boundaries and

PAGBs after tempering at 790°C for 4 hours in 9Cr steel with 0.07 wt.% nitrogen [26].

During creep at 650°C, the agglomeration and coarsening rates of MX carbonitrides in

steel with 0.07 wt.% nitrogen were greater than those in steel with 0.05 wt.% nitrogen. It

is believed that there should be an optimal content of nitrogen relative to other nitride

forming elements such as boron [22].

24
(ppm)
Figure 2.15: Time to rupture and minimum creep rate of 9Cr-3W-3Co-0.2V-0.05Nb-
0.08C steel with 140ppm B at 650°C and 120Mpa as a function of nitrogen concentration
[18].

25
2.1.3.6 Molybdenum

Molybdenum is a ferrite stabilizer and strong carbide former. It was believed that

for a 2.25Cr-Mo-0.25V steel, molybdenum increased the creep resistance of steel when it

was present either in solid solution or as carbide precipitate [29]. It was found that for a

3Cr-Mo steel, the rupture time tested at 538°C and different stress levels increased with

increasing of Mo from 0.8 wt.% to 1.6 wt.% [29].

In a recent research conducted by Hald, it was proposed that solid solution

strengthening from molybdenum had no significant effect on long-term microstructure

stability of 9-12Cr steels, because at 600°C the solute molybdenum atoms diffused at a

rate similar to the dislocation velocity [3]. Laves phases (Fe,Cr)2Mo or (Fe,Cr)2(Mo,W)

would form during long-term aging process. Mo is also one of the strong M23C6 carbide

formers.

26
2.1.3.7 Tungsten

Tungsten is a ferrite stabilizer. It was believed that the replacement of part of

molybdenum by tungsten in NF616 (P92) steel improved the creep rupture strength by

solid solution strengthening provided by tungsten [12]. It was reported that for a 9Cr-

0.5Mo-W steel both tensile strength and creep rupture strength increased with increasing

tungsten content [28]. As shown in Figure 2.16, at 600°C and 700°C, the creep rupture

strength of a 9Cr-0.5Mo-W steel increased with increasing the tungsten content. It was

believed that this effect was due to solid solution hardening by tungsten. However, as

shown in Figure 2.17, the impact toughness decreased with increasing tungsten content in

both tempered and aged conditions. Considering the creep rupture strength and

toughness, the optimum tungsten content was determined to be around 2 wt.%.

Recently Hald suggested that solid solution strengthening from tungsten had no

significant effect on long-term microstructure stability of 9-12Cr steels [3]. It was found

that after aging at 600°C for 50000 hours, tungsten in P92 steel produced stable and finer

Laves phase particles than those in P91 steel, thus improving the creep rupture strength of

P92 steel. This was explained by the W-rich Laves phase (Fe,Cr)2(Mo,W) having a

higher dissolution temperature (720°C) than Mo-rich Laves phase (Fe,Cr)2Mo (650°C).

Aging temperature (600°C) is close to the dissolution temperature of Mo-rich Laves

phase (650°C), so nucleation of Mo-rich Laves phase became very difficult and only a

few particles nucleated then grown into extended large particles. In P92 steel, loss of

tungsten from solid solution through Laves phase precipitating also improved the stability

of M23C6 carbides, because tungsten is also one of the strong M23C6 carbide formers.

27
Figure 2.16: Creep rupture strength of 9Cr-0.5Mo-W steels tested at 600°C and 700°C
[28].

Figure 2.17: Tensile properties of 9Cr-0.5Mo-W steels tested at room temperature [28].
28
2.1.3.8 Molybdenum versus Tungsten

As discussed above both molybdenum and tungsten were believed to be solid

solution strengthening elements [12] [29], later research found both of them were strong

M23C6 carbides formers and during the aging process they were strong Laves phases

formers [3]. So the combination effect of molybdenum versus tungsten was of interest.

As shown in Figure 2.18(a), increasing the tungsten ratio while keeping the Mo

equivalent (Mo+0.5W) at 1.5 wt.% was the most efficient for creep strengthening [30].

Result of further research on 12Cr turbine steels is shown in Figure 2.18(b). The content

of tungsten and molybdenum were changed by increasing tungsten and decreasing

molybdenum on the basis of 1 wt.% W and 1 wt.% Mo in 12Cr steel. The time to rupture

was found to be longest at the combination of 1.8 wt.% W and 0.7 wt.% Mo [31]. The

effects of molybdenum and tungsten concentrations on creep rupture strength at 650°C

for 105 hours and the Charpy impact energy are shown in Figure 2.14 [32]. It was

empirically shown that creep strength achieved peak when the Mo equivalent (Mo+0.5W)

was set at 1.5 wt.%. It was found that increasing tungsten content improved creep

strength while reduced ductility and toughness. The line in Figure 2.19 represents Mo

equivalent of 1.5 wt.%. Consequently, for particular composition of a 9-12Cr steel there

should be an optimal balance of tungsten and molybdenum content.

29
(b)
(a)

Figure 2.18: (a) Effect of Mo+W on creep rupture strength of 12% Cr steels [27]; (b) Effect of
Mo and W on 105h creep rupture strength at 650°C and Charpy absorbed energy at 20°C of 12%
Cr turbine steels (0.13C-0.05Si-0.5Mn-11.2Cr-0.8Ni-0.2V-0.5Nb-0.05N-Mo-W) [31].

Figure 2.19: Effect of W+Mo on creep rupture strength and Charpy absorbed energy in
12Cr turbine steels [32].

30
2.1.3.9 Copper

Copper is austenite stabilizer, high copper was added to HCM12 for the first time

in order to suppress the formation of δ-ferrite and without reducing the creep strength [5].

Addition of copper enhanced the resistance to HAZ softening without deterioration of

creep rupture strength and ductility [10]. As shown in Figure 2.12, copper did not

decrease the creep rupture strength at 600°C for 105 hours up to 2 wt.% [5]. However, it

was found that for copper contents up to 2 wt.%, after creep rupture at 650°C and 98Mpa

load for 5306 hours, fine copper phase precipitated on the lath boundaries and within the

lath. These precipitates are thought not to be effective in creep strength. It was proposed

that guarantee the high creep strength, copper should be less than 2 wt.% [5].

Copper precipitation in a P122 steel with 0.87 wt% copper was researched by

Hattestrand and Andren. They reported a dense distribution of copper precipitates formed

during tempering at 770°C [27]. During aging at 600°C or 650°C, copper precipitates

coarsened rapidly were believed to be less effective for long-term creep strength than

M23C6 and VN precipitates. However, compared to P92 steel a finer distribution of Laves

phase was observed in P122 steel. This was believed to be an effect of accelerated

nucleation, because of copper precipitates working as suitable nucleation site for Laves

phase [27].

31
2.1.3.10 Vanadium and Niobium

Both vanadium and niobium are strong MX carbonitrides forms, and MX

carbonitrides are extremely stable against coarsening at elevated temperature (600°C) [3].

The combination effect of vanadium versus niobium on 104h creep rupture strength at

600°C and 650°C in NF616 (P92) with low carbon (0.05 wt.%) was researched by

Nippon Steel [33]. As shown in Figure 2.20, the optimized creep strength was obtained

when niobium was round 0.05 wt.% and the optimal vanadium content was around from

0.10 wt.% to 0.25 wt.%. A similar result was proposed for 12Cr steels, with optimal

contents for vanadium and niobium of approximately 0.05 wt.% and 0.20 wt.% [30].

(a) (b)

Figure 2.20: (a) Effect of V and Nb on 104h creep rupture strength of NF616 at 600°C; (2)
Effect of V and Nb on 104h creep rupture strength of NF616 at 650°C [30].

32
A recent study about effects of vanadium and niobium on precipitation behavior

and mechanical properties of a 0.11C-0.7Mn-10.2Cr-1.2Mo-0.05N-V-Nb was done by

Onizawa et al. [34]. Figures 2.21 shows higher strength was obtained with increasing

vanadium and niobium contents, but the impact properties degrade with the increase of

vanadium and niobium contents. The ductility also degraded with increasing vanadium

and niobium contents [34]. As shown in Figure 2.22 and 2.23, the creep rupture time

became longer with increasing vanadium and niobium contents, however the effect of

niobium tended to be saturated at about 0.01 wt.%. It was observed that fine MX

carbonitrides V(C,N) and Nb(C,N) formed in this 10Cr steel, the number of MX

carbonitrides increased with the contents of vanadium and niobium. During aging at

873K for 6000 hours MX carbonitrides did not grow and not any new precipitates were

observed.

Figure 2.21: (a) Influence of V and Nb on tensile properties ultimate tensile strength at
823K; (b) Influence of V and Nb on ductile-brittle transition temperature after aging at
873K for 6000h [34].

33
Figure 2.22: Influence of Nb on creep rupture time at 873K [34].

Figure 2.23: Influence of V on creep rupture time at 873K [34].

34
2.1.3.11 Cobalt

The composition range of cobalt is not defined in ASTM specification for P91,

P92 and P122. Cobalt was added to a 12Cr-W steel in order to suppress the formation of

δ-ferrite like copper [36]. It was reported that δ-ferrite was entirely eliminated when 1.05

wt.% cobalt was added. I was also found that cobalt improved the creep rupture strength

at 650°C for 104 hours; however the influence of cobalt was saturated when concentration

of cobalt was around 1 wt.% [36][37].

2.1.3.12 Tantalum

For nuclear fusion application tantalum was added into Cr steels to replace

niobium in order to meet reduced-activation criteria [38] [39]. It was proposed that

tantalum increased resistance to irradiation-induced embrittlement of a 9Cr-2WVTa steel,

because tantalum in the solution caused either an increase in fracture stress or a change in

flow behavior [39]. Tantalum is a MX carbonitride former; it was also found that

tantalum entered Z-phase as a minor element [40]. Recent research by Y. Wang et al.

found a Ta-alloyed11%Cr steel exhibited a distinctively higher creep strength at 650°C

than the advanced 9%Cr steel and piping steel P92 [41]. It was believed this was due to

enhanced strengthening by fine precipitates, in particular tantalum containing MX

cabonitride precipitates that were found by high resolution TEM.

35
2.1.3.13 Manganese and Silicon

Manganese is an austenite stabilizer that could suppress the formation of δ-ferrite

and has a significant influence on A1 transformation temperature as nickel [22].

Manganese and nickel could impair high temperature stability of the ferrite structure by

decreasing A1 transformation temperature. Mi+Ni contents were believed to be a

significant factor for the A1 temperature in steels. It was believed that the steels of higher

A1 temperature may have better creep rupture strength; copper and cobalt also could

suppress the formation of δ-ferrite but decrease A1 temperature less than manganese and

nickel [44]. Compared to manganese and nickel, copper and cobalt are better alloying

elements. Silicon is a ferrite stabilizer and silicon decreased toughness of steel by

promoting Laves phase precipitation. Reduction of both manganese and silicon could

improve creep strength of steel [22].

36
2.2 Phase Transformation Analysis Techniques

2.2.1 Differential Thermal Analysis (DTA)

As shown in Figure 2.24(a), Differential Thermal Analysis (DTA) requires one

test sample and one reference sample [48]. These two samples are put in the same

environment and undergo identical heat flux. Thermal couples are used to record the

instant temperature of samples. During the thermal cycle phase transformation does not

take place in the inert reference sample. The heat absorption and emission relative with

phase transformation in the testing sample can be detected by comparing the instant

temperature records of testing sample and reference sample. As shown in Figure 2.24(b),

this is a difference between the instant temperatures of the test sample and the reference

sample as a function of the temperature [49]. A local deviation (peak) in the curve

represents a phase transformation. The start and the end of this deviation correspond to

the beginning and the finish of a phase transformation.

(a) (b)

Figure 2.24: (a) Schematic of DTA installation [48]; (b) Typical DTA curves of low
alloy Cr-Mo steel a- hot rolled and b- as quenched [49].
37
2.2.2 Differential Scanning Calorimetry (DSC)

A Differential Scanning Calorimetry operation system is shown in Figure 2.25(a).

The test and the reference sample are put on material with high heat conductivity and

undergo identical heat flux [50]. Because of high heat conductivity of sample holder, the

difference between instant temperatures of test sample and reference sample relative to

phase transformation would induce a heat flux between them through sample holder. As

shown in Figure 2.25(b), a peak in the heat flux record represents a phase transformation.

a)

b)

Figure 2.25: (a) Schematic of DSC installation; (b) Typical DSC result. [50]

38
2.2.3 Dilatometry

In 1982, Bhadeshia detected the difference between linear thermal expansion

coefficients of ferrite and austenite in steels [51]. He studied three steel samples, two

designed alloys Fe-0.39C-2.05Si-4.08Ni and Fe-0.22C-2.03Si-3.0Mo, and one

commercial steel 300M. Samples were machined in shape of 3.2mm outer diameter,

1.5mm inner diameter and 20mm length cylinder rod. The linear thermal expansion was

detected by a length transducer measuring the diameter change of cylinder rod samples.

Thermal expansion of ferrite and austenite were studied at 25°C-600°C and 800°C-990°C

respectively. As shown in Figure 2.26, for these three steels linear thermal expansion

coefficient of austenite was higher than that of ferrite. Composition could influence the

linear thermal expansion coefficient of ferrite and austenite.

In 1989, after studied the isothermal transformation of austenite in a Fe-0.3C-

4.08Cr steel at 420°C, Takahashi and Bhadeshia proposed that the relative diameter

change was proportional to the volume fraction of ferrite, till at least volume fraction of

ferrite arrived 0.7 [52].

Figure 2.26: Linear thermal expansion coefficient of ferrite and austenite [51].

39
The application of dilatometry to study the phase transformations in

hypoeutectoid carbon and low alloy steel is defined in ASTM A1033-04. If there is no

phase transformation, relative diameter change (linear strain) should be proportional to

temperature during continuous heating or cooling process. If phase transformation

happens, because of difference in linear thermal expansion coefficients of different

phases, the linear feature of linear strain curve as a function of temperature would be

changed. As shown in Figure 2.27, a change in the linearity of the strain vs. temperature

curve represents a volume change that is related to a certain phase transformation.

Figure 2.27: Typical Dilatometry curve of steel [ASTM A 1033-04].

40
2.2.4 Single Sensor Differential Thermal Analysis (SS-DTA)

Single Sensor Differential Thermal Analysis (SS-DTA) has been recently

developed by Alexandrov and Lippold [53].The principle of SS-DTA is via detecting the

latent heat of phase transformation to determine the phase transformation starting and

finishing temperatures. Compared to DTA and DSC, SS-DTA does not need a reference

sample. SS-DTA uses a reference temperature curve that is calculated based on actual

thermal history of the tested material. As shown in Figure 2.28, the tested sample is put in

argon atmosphere in a resistive heated convection furnace. Type K thermocouples are

used to record the instant sample temperature [54]. In Figure 2.29(a), the blue curve

represents the temperature history of the test sample and red curve represents the

calculated reference curve. As shown in Figure 2.29(b), the difference between sample

and the reference cooling curves is plotted as a function of sample temperature, deviation

peak in this curve represents phase transformation; the start and the end of the deviation

peak indicates the beginning and finish of corresponding phase transformation [54].

41
Sample Inconel Lid
tube

Compression
Ar out fitting

Ar in
Ar in

Thermocouple
Furnace tube Ceramic Gasket
insulator

Figure 2.28: Schematic of SS-DTA installation [54].

(a) (b)

Figure 2.29: SS-DTA result of steel HSLA-65: (a) sample cooling curve and reference
curve; (b) difference between sample cooling curve and reference curve as a function of
sample temperature [54].

42
2.2.5 The Electrothermomechanical Tester (ETMT)

As shown in Figure 2.30, the Electrothermomechanical Tester (ETMT) at Ohio

State University was developed by Instron and NPL [55]. ETMT could detect the change

in resistance of metal samples relative to phase transformation thus inspect the phase

transformation process in metal samples. ETMT heated the sample in size of 40*2*1mm

by resistive heating, thermal couple was applied to measure the sample temperature. At

the temperature when resistance started to change indicates the beginning of phase

transformation in sample.

Figure 2.30: Image of ETMT [55]

43
2.3 Predicting Methods for Critical Temperatures

2.3.1 Predictive Formulae for Ms and Mf Temperatures

The development of predictive formulae for Ms temperature in steels is shown in

Table 2.2. In 1944 [56], by studying 17 designed steels made in the lab Payson and

Savage developed the first formula for Ms temperature as a linear function of

compositions. At that time the Ms temperature was determined by optical microscopy

study of quenched and then tempered steel samples [56]. In order to achieve better

agreement, Carapella applied “multiplying-factor-principle” to Payson’s formula and

developed a formula for Ms temperature by multiplying composition values [57]. In 1946

[58], Rowland and Lyle adjusted the coefficient of carbon in Payson’s function from -

316.7 to -333.3, so the deviation between the calculated value (by Payson’s formula) and

experimental value of Ms temperature in nine commercial steels became smaller. In 1946

[59], by collecting all available published data, Grange and Stewart developed a new

predictive formula for Ms temperature. The calculated Ms temperature were all within 6.7

°C of the measured value of 13 commercial low-alloy steels. At the same time, Payson’s

formula was revised by Nehrenberg, coefficient of C was changed from -316.7 to -300

[57]. In 1956 [60], Steven and Haynes studied 59 commercial steels and several steel

samples made in the lab. The method of least squares was used to analyze measured Ms

temperatures. Constant of Ms formula was adjusted to 561.4, coefficient of C, Cr and Mo

also were all modified. Calculated Ms temperature by Steven and Haynes’s formula had

much smaller deviation from measured values in steels with less than 0.25 wt.% carbon.

44
In 1965 [1], Andrew studied available data of 184 commercial steels including

the 59 from Steven and Hayne’s study. Both a linear formula and a product formula were

developed. For linear formula, the coefficient of Mo was adjusted to much lower,

constant and coefficient of C were also modified. The possible influence of second order

terms were taken into account in the product formula. It was found that Cr content was

influenced by C interaction especially when Cr was above 2 wt.%. In 1982 [61],

C.Y.Kung and J.J.Rayment evaluated the validity of all available predictive formula for

Ms temperature; it was proposed that cobalt should be included in formula for Ms

temperature and coefficient of cobalt was 10. The term of Co was added to each formula

in Table 2.2. In 1992 [62], by through literature survey, Jicheng Zhao developed

predictive formulae for Ms of both lath martensite and twinned martensite. Zhao’s

formula had the largest applicable composition range. In 1994 [63], Beres and Irmer

developed a new formula for Ms temperature, as shown in Figure 2.31. The Ms

temperature was influenced by composition of Mn, Ni, Mo and value of term Msc and Cra.

The term of Msc and Cra were calculated by defined function, and coefficient X, Y and Z

varied with composition range as indicated in Figure 2.31. The applicable composition

ranges for all these formulae are shown in Table 2.3.

45
Author Formulae

Payson and Ms(˚C)=498.9-316.7C-33.3Mn-27.8Cr-16.7Ni-11.1Si-


Savage(1944) 11.1Mo-11.1W (+10Co)
Carapella Ms(˚C)=496*(1-0.62C)(1-0.092Mn)(1-0.07Cr)(1-
(1944) 0.045Ni) (1-0.0033Si)(1-0.029Mo)(1-0.018W)
(1+0.012Co)
Rowland and Ms(˚C)=498.9-333.3C-33.3Mn-27.8Cr-16.7Ni-11.1Si-
Lyle (1946) 11.1Mo-11.1W (+10Co)
Grange and Ms(˚C)=537.8 -361.1C-38.9(Mn+Cr)-19.4Ni-27.8Mo
Stewart (1946) (+10Co)
Nehrenberg Ms(˚C)=498.9-300C-33.3Mn-22.2Cr-16.7Ni-11.1Si-
(1946) 11.1Mo (+10Co)
Steven and Ms(˚C)=561.1-473.9C-33.3Mn-16.7(Cr+Ni)-21.1Mo
Haynes (1956) (+10Co)
Andrew (1965) Ms(˚C)=539-423C-30.4Mn-17.7Ni-12.1Cr-7.5Mo
(+10Co)
2
Ms(˚C)=512-453C-16.9Ni+15Cr-9.5Mo+217C -
71.5Mn*C-67.6Cr*C (+10Co)
Zhao (1992) MsLM (˚C)=540-356.25C-206.64N-24.65Ni+1.36Ni2-
0.0384Ni3-17.82Cr+1.42Cr2-47.59Mn+2.25Mn2-
0.0415Mn3+17.50Mo +21.87Co-0.468Co2+
0.00296Co3- 16.52Cu-17.66Ru

MsTM (˚C)=420-208.33C-72.65N-43.46N2-16.08Ni
+0.7817Ni2-0.02464Ni3-2.473Cr -
33.428Mn+1.296Mn2-0.02167Mn3
+30.00Mo+12.86Co-0.2654Co2+ 0.001547Co3 -
7.18Cu-16.28Ru+1.72Ru2-0.08117Ru3
Table 2.2: Development of Ms temperature predictive formulae [61]

46
Ms(˚C)=Msc-YMn-XNi-ZCra-7.5Mo
Msc (˚C)= 454-210C+4.2/C 0.03≤C% ≤0.35
Msc (˚C)= 332-190C+40/C 0.35<C% ≤1.3
Msc (˚C)= 116 1.3≤C% ≤2.3
Cra=Cr+1.5Si+W+V+Al+0.5Nb+0.5Ta+2Ti

Content of elements (wt.%) Coefficients


Group
C, Cr,Si Ni X Y Z
>5 21 10.5 16.8
0.03≤C<0.5 and
1 1.4~5 1.6*Ni+65/Ni 10.5 16.8
(Cr+1.5Si)>6
<1.4 27 7.8 9.5
0.03≤C<2.3 and Mn≥1.75 Mn<1.75
>5 15.6 2.1
(Cr+1.5Si)≤6 31.2 7.8
2 or 2~5 1.2*Ni+48.7/Ni 31.2 7.8 2.1
0.5≤C<2.3 and
(Cr+1.5Si)>6 <2 27 31.2 7.8 2.1

Figure 2.31: Formula of Ms developed by Beres and Irmer(1994) [63].

Steven and Haynes measured the amount of martensite formed when sample was

quenched to designed temperatures [60]. It was reported that the first 70% martensite

formed rapidly when quench temperature decreasing from Ms; however at lower quench

temperatures formation of martensite became much slower. They found the following

relationships (Mx means x% martensite formed):

M10(˚C) = Ms-10 ±3
M50(˚C) = Ms-47 ±9
M90(˚C) = Ms-103 ±12
Mf (˚C) = Ms-215 ±15

47
Payson Rowland Grange Nehrenberg Steven Andrew Zhao Beres
and and and (1946) and (1965) (1992) and
Savage Lyle Stewart Haynes Irmer
(1944) (1946) (1946) (1956) (1994)
C <1.0 <1.0 <0.9 0.17-1.28 0.10- 0.110- <2 0.03-
0.55 0.600 2.30
Mn 1.02- 0.45- <7.0 0.29-1.21 0.20- 0.040- <27 -
4.87 0.82 1.70 4.870
Si <1.0 0.21- - 0.15-1.89 - 0.110- - All
0.31 1.890 range
Cr 0.98 0.02- <1.5 0.05-8.81 <3.50 <4.610 <10 All
-4.61 1.50 range
Ni 1.16- 0.08- <14.0 0.10-3.41 <5.0 <5.040 <34 All
4.83 3.99 range
Mo 1.05- 0.02- <1.0 0.03-0.33 <1.00 <5.400 <2.5 -
5.40 0.28
W 1.04- - - - - - - -
4.78
N - - - - - - <3 -
Co - - - - - - <60 -
Cu - - - - - - <7 -
Ru - - - - - - <20 -
Table 2.3: Applicable composition range of formulae for Ms (wt.%) [1][56]~[63].

48
2.3.2 Predictive Formulae for Bs and Bf Temperatures

Available formulae of Bs and Bf temperatures were shown in Table 2.4. The first

formula for Bs temperature was developed by Steven and Haynes; they also developed

formulae for B50 and Bf [60]. In 1992 [62], via through literature survey, Jicheng Zhao

developed a formula for Bs temperature. The applicable composition ranges for these two

formulae are shown in Table 2.5. In 2000 [64], Z. Zhao et al. developed a formula for Bs

temperature; carbon was eliminated from this formula. It was reported that for 84 low

alloy steels this formula achieved better agreement than Steven and Haynes’ formula. It

was also proposed by Z. Zhao et al. that Bs temperature was approximately 50-55˚C lower

than B0 (the theoretical upper temperature limit) temperature. B0 was derived from the

intersection of thermodynamic equilibrium curves of austenite to ferrite transformation

and austenite to cementite transformation.

49
Author Formulae

Steven and Bs(˚C)=830-270C-90Mn-37Ni-70Cr-83Mo


Haynes (1956) B50(˚C)= Bs-60 Bf(˚C)= Bs-120
Zhao(1991) Bs(˚C)=720-585.63C-126.60C2-63.34Ni+6.06Ni2-
0.232Ni3 -31.66Cr+2.17Cr2-91.68Mn+7.82Mn2-
0.3378Mn3-42.37Mo+9.16Co-0.1255Co
+0.000284Co3-36.02Cu-46.15Ru
Z.Zhao et Bs(˚C)=630-45Mn-45V-35Si-30Cr-25Mo-20Ni-15W
al.(2000)
Table 2.4: Formulae for Bs and Bf [60] [62] [64].

C Mn Si Ni Cr Mo N Cu Co Ru
Steven
and 0.10- 0.20-
- <5.0 <3.5 <1.0 - - - -
Haynes 0.55 1.70
(1956)
Zhao
<0.8 <10 - <20 <10 <2.5 <0.6 <7 <20 <5
(1991)
Z.Zhao
et al. For commercial low alloy steels
(2000)
Table 2.5: Applicable composition range of formulae for Bs and Bf (wt.%) [60] [62] [64].

50
2.3.3 Predictive Formulae for Ac1 and Ac3 Temperatures

As shown in Table 2.6, the only available predictive formulae for Ac1 and Ac3

temperatures in steels were developed by Andrew in 1965 [1]. Andrew collected

available data of 196 Ac1 temperatures and 155 Ac3 temperatures. The applicable

composition ranges for these two formulae were shown in Table 2.7.

Formulae
Ac1 Ac1(˚C)=723-10.7Mn-
16.9Ni+29.1Si+16.9Cr+290As+6.38W
Ac3 Ac3(˚C)=910-230√C-
15.2Ni+44.7Si+104V+31.5Mo+13.1W

Table 2.6: Formulae for Ac1 and Ac3 [1].

C Mn Si Ni Cr Mo N V W
Ac1 0.08- 0.04- 0.06- 0.00- 0.00- 0.00- 0.00- 0.00- 0.08-
1.43 1.98 1.78 5.00 4.48 4.48 2.51 4.10 1.43
Ac3 0.08- 0.04- 0.09- 0.00- 0.00- 0.00- 0.00- 0.00- 0.08-
0.59 1.98 1.78 5.00 4.48 1.05 0.70 4.10 0.59
Table 2.7: Applicable composition range of formulae for Ac1 and Ac3 (wt.%) [1].

51
Chapter 3 OBJECTIVES

The objective of this study was to develop predictive formulae for the A1

temperature in P91, P92 and P122 steels as a function of steel composition. The research

steps were designed as follows:

1. Screen out the alloying elements with a significant effect on the A1 temperature in

P91 steel using a model based DOE. The JMP8 software was utilized to design

the modeling experiments and the thermodynamic software JMat-Pro was used to

predict the A1 temperature.

2. Develop a predictive formula for the A1 temperature in P91 steel as a function of

the content of metallic alloying elements using a measurement based DOE. The

JMP8 software was utilized to design the experiments and the SS DTA technique

was used to measure the A1 temperature in laboratory heats of P91 steel.

3. Study the effect of C, Si, and N on the A1 temperature in P91 steel.

4. Study the effect of Mo and W on the A1 temperature in P92 steel.

5. Study the effect of Cu on the A1 temperature in P122 steel.

6. Develop the predictive formulae for the A1 temperature in P91 and P92 steels.

7. Validation of the formulae for predicting the A1 temperature in P91 and P92

steels.

52
Chapter 4 MATERIALS AND EXPERIMENTAL PROCEDURES

4.1 Design of Experiments

4.1.1 Model Based DOE

The chemical compositions of P91, P92 and P122 steels specified by ASTM

A335 are shown in Table 4.1. In order to screen out alloying elements with significant

effect on predicted A1 temperature in P91 steel, DOE software JMP8 was used to fulfill

effect screening. Samples with various compositions inside the ASTM specification

range were designed using JMP8, and their A1 temperatures were calculated by the

thermodynamic simulation software JMat-Pro. By analyzing the data, the alloying

elements with significant influence on the predicted A1 temperature in P91 steel were

determined. The results of the JMat-Pro based DOE were used to select the significant

elements for further experimental study. The structure of the model based DOE used in

this study is shown in Figure 4.1

The JMP software is designed by SAS Institute to perform statistical analysis.

JMP8 could also be used to design experiments. As shown in Figure 4.2, at the JMP8

starter menu, DOE was chosen to start experimental design. It was assumed that the

alloying elements have quadratic relation with the predicted A1 temperature. Custom

Design was chosen so that both linear and quadratic terms could be defined. As shown in

Figure 4.3, at first, the response of model based DOE was ready to be defined. Here the

53
response is JMat-Pro predicted A1 temperature and it was defined as A1 in JMP8. More

than one response could be defined for one experimental deign. As shown in Figure 4.4,

factors for model based DOE were designed after response was fixed. Here the factors

are the contents of alloying elements in P91 steel. Because composition is numeric data,

composition of each alloying element was defined as continuous factor. The composition

range was also defined according to ASTM specification for P91 steels as shown in Table

4.1. All of the alloying elements were defined as factors except for P, S and Al. The

effect of these elements on predicted A1 temperature is negligible because their

concentrations are extremely low.

Figure 4.1: Structure of the model based DOE used in this study.

54
Figure 4.2: JMP8 starter menu and DOE features.

Figure 4.3: Defining the response for model based DOE in JMP8.

55
Figure 4.4: Defining the factors for model based DOE in JMP8.

In order to find out all possible relations between the steel composition and the

JMat-Pro predicted A1 temperature, linear term, quadratic term and second order

interaction term were all defined for model based DOE model, as shown in Figure 4.5.

Figure 4.6 shows how to define number of center points and replicates for model based

DOE. Because the response of this experimental design is software JMat-Pro calculated

A1 temperature, there would be no deviation between replicated runs with the same

composition. This is why no replicates and one center point were defined. Under the

Custom Design of DOE features in JMP8, response, factors, model, number of center

points and replicates were defined for model based DOE step by step. As shown in Table

4.2 and 4.3, the model based DOE table was designed using JMP8. There were totally 65

runs, which means 65 P91 steels with differnent composition were designed using JMP8.

The A1 temperature in all 65 compositions of P91 steel were predicted using JMat-Pro

and were included in the response column for A1, as shown in Tables 4.2 and 4.3.

56
Figure 4.5: Defining the model effects for model based DOE in JMP8.

57
Figure 4.6: Defining the number of center points and replicates for model based DOE in
JMP8.

Element P91 P92 P122


C 0.08-0.12 0.07-0.13 0.07-0.14
Mn 0.30-0.60 0.30-0.60 0.30-0.70
Si 0.20-0.50 <0.50 <0.50
S <0.010 <0.010 <0.010
P <0.020 <0.020 <0.020
Cr 8.00-9.50 8.50-9.50 10.00-12.50
Ni <0.40 <0.40 <0.50
Cu - - 0.30-1.70
Mo 0.85-1.05 0.30-0.60 0.25-0.60
W - 1.50-2.00 1.50-2.50
V 0.18-0.25 0.15-0.25 0.15-0.30
Nb 0.06-0.10 0.04-0.09 0.04-0.10
N 0.030-0.070 0.030-0.070 0.040-0.10
B - 10-60ppm <0.005
Al <0.040 <0.040 <0.040
Table 4.1: ASTM A335 specification for P91, P92 and P122 steels (wt.%).

58
Run C Mn Si Cr Ni Mo V Nb N A1/°C
1 0.08 0.3 0.2 8 0.1 1.05 0.18 0.1 0.07 816
2 0.08 0.3 0.35 9.5 0.1 0.85 0.18 0.06 0.07 819
3 0.08 0.6 0.5 9.5 0.4 0.85 0.25 0.1 0.03 791
4 0.1 0.6 0.5 8 0.1 1.05 0.18 0.06 0.07 804
5 0.08 0.3 0.5 9.5 0.4 0.85 0.18 0.1 0.07 801
6 0.12 0.3 0.5 9.5 0.1 0.85 0.18 0.06 0.03 825
7 0.12 0.3 0.2 9.5 0.1 1.05 0.18 0.1 0.03 823
8 0.08 0.3 0.5 8 0.1 1.05 0.25 0.1 0.03 830
9 0.12 0.6 0.2 8 0.4 0.85 0.18 0.1 0.03 777
10 0.12 0.3 0.2 9.5 0.4 1.05 0.18 0.06 0.07 796
11 0.12 0.45 0.5 8 0.1 1.05 0.18 0.1 0.03 816
12 0.08 0.3 0.2 9.5 0.1 1.05 0.25 0.06 0.03 830
13 0.12 0.3 0.2 8 0.1 0.95 0.25 0.1 0.05 816
14 0.08 0.3 0.5 9.5 0.1 1.05 0.18 0.1 0.03 834
15 0.1 0.6 0.35 8 0.1 0.95 0.215 0.1 0.03 805
16 0.12 0.6 0.2 9.5 0.4 0.85 0.25 0.1 0.07 777
17 0.1 0.3 0.2 8 0.1 0.85 0.25 0.06 0.07 812
18 0.08 0.3 0.5 8 0.4 1.05 0.18 0.06 0.07 802
19 0.12 0.6 0.5 8 0.25 0.95 0.25 0.06 0.07 793
20 0.12 0.3 0.2 8 0.4 1.05 0.25 0.06 0.03 800
21 0.08 0.6 0.2 8 0.1 0.85 0.25 0.1 0.07 798
22 0.08 0.6 0.5 8 0.25 0.85 0.18 0.08 0.03 796
23 0.08 0.3 0.5 8 0.4 0.95 0.25 0.06 0.03 807
24 0.12 0.6 0.5 9.5 0.4 0.85 0.18 0.06 0.07 780
25 0.12 0.6 0.2 8 0.4 1.05 0.18 0.1 0.07 777
26 0.08 0.6 0.2 8 0.1 1.05 0.25 0.06 0.07 801
27 0.12 0.6 0.5 9.5 0.1 1.05 0.18 0.1 0.05 810
28 0.12 0.45 0.2 9.5 0.1 0.85 0.25 0.06 0.03 813
29 0.08 0.6 0.5 9.5 0.1 0.95 0.25 0.06 0.03 817
30 0.12 0.6 0.35 9.5 0.4 0.95 0.18 0.08 0.03 784
31 0.12 0.6 0.5 8 0.4 1.05 0.215 0.06 0.03 787
32 0.12 0.3 0.35 8 0.4 0.85 0.18 0.06 0.05 795
Table 4.2: The first part of model based DOE table with JMat-Pro predicted A1
temperature values (wt.%).

59
Run C Mn Si Cr Ni Mo V Nb N A1/°C
33 0.1 0.45 0.5 9.5 0.1 1.05 0.25 0.1 0.07 821
34 0.12 0.3 0.5 9.5 0.1 1.05 0.25 0.06 0.03 833
35 0.08 0.3 0.2 9.5 0.4 0.85 0.18 0.06 0.03 801
36 0.08 0.3 0.2 8 0.4 1.05 0.18 0.1 0.03 801
37 0.08 0.6 0.2 9.5 0.1 0.85 0.18 0.1 0.03 806
38 0.12 0.3 0.5 8 0.4 0.85 0.25 0.1 0.03 803
39 0.1 0.45 0.2 8.75 0.4 0.95 0.25 0.08 0.05 790
40 0.08 0.6 0.2 9.5 0.4 1.05 0.18 0.1 0.07 782
41 0.12 0.6 0.2 8.75 0.1 1.05 0.18 0.06 0.03 803
42 0.08 0.3 0.5 9.5 0.1 0.85 0.25 0.08 0.07 826
43 0.08 0.45 0.35 9.5 0.25 1.05 0.18 0.06 0.03 810
44 0.1 0.6 0.2 9.5 0.25 0.85 0.215 0.06 0.05 791
45 0.12 0.3 0.35 8 0.4 1.05 0.25 0.1 0.07 799
46 0.12 0.6 0.2 8 0.1 0.85 0.18 0.08 0.07 793
47 0.08 0.6 0.2 8 0.4 0.85 0.25 0.06 0.03 781
48 0.08 0.3 0.35 8.75 0.25 0.85 0.25 0.1 0.03 815
49 0.12 0.3 0.2 9.5 0.25 0.85 0.18 0.1 0.07 803
50 0.12 0.6 0.5 8.75 0.1 0.85 0.25 0.1 0.03 808
51 0.12 0.3 0.5 8 0.1 1.05 0.215 0.08 0.07 821
52 0.08 0.3 0.2 8 0.1 0.85 0.18 0.06 0.03 816
53 0.08 0.3 0.2 8 0.4 0.85 0.215 0.08 0.07 793
54 0.08 0.6 0.5 9.5 0.4 1.05 0.25 0.06 0.07 789
55 0.12 0.3 0.5 9.5 0.4 0.85 0.25 0.06 0.07 800
56 0.08 0.6 0.5 8.75 0.1 0.95 0.215 0.1 0.07 807
57 0.1 0.3 0.5 8 0.1 0.85 0.18 0.1 0.07 817
58 0.12 0.6 0.2 9.5 0.25 1.05 0.25 0.1 0.03 798
59 0.08 0.6 0.5 8 0.4 1.05 0.25 0.1 0.05 789
60 0.12 0.3 0.5 8.75 0.4 1.05 0.18 0.1 0.03 806
61 0.08 0.6 0.2 8 0.4 0.95 0.18 0.06 0.07 776
62 0.08 0.45 0.5 8 0.1 0.85 0.215 0.06 0.05 813
63 0.08 0.3 0.2 9.5 0.4 1.05 0.25 0.1 0.07 803
64 0.12 0.6 0.2 9.5 0.1 1.05 0.25 0.06 0.07 803
65 0.1 0.45 0.35 8.75 0.25 0.95 0.215 0.08 0.05 802
Table 4.3: The second part of model based DOE table with JMat-Pro predicted A1
temperature values (wt.%).

60
With the JMat-Pro prodicted A1 temperatures, the model based DOE data were

ready to be analyzed using JMP8. As shown in Figure 4.7, a model was tailored for

model based DOE data in the model fitting platform. A1 was defined as the role variable

of the model. The model effects were the same as we defined in Figure 4.5, including all

linear terms, quadratic terms and second order interaction terms. The JMat-Pro predicted

A1 temperature data was modeled by a standard least square approach. Effects screening

was defined as model fitting emphasis, thus factors with significant effects will be

showed in a scaled parameter report with a graph and the prediction profiler.

Figure 4.7: The model fitting platform for model based DOE in JMP8.
61
4.1.2 Measurement based DOE and separate experimental study

The alloying elements with significant influence on JMat-Pro predicted A1

temperatures in P91 steel were screened out using the JMP8 software. These significant

elements were selected as factors in our measurement based DOE. Because we could not

control the concentration of C, N and Si as accurate as of the metallic elements, only

metallic elements with significant effect on predicted A1 temperatures were defined as the

factors of DOE model in this study. Additionally, we investigated the influence of C, N

and Si on A1 temperatures in P91 steel, the effect of Mo and W on A1 temperatures in

P92 steel and the effect of Cu on A1 temperatures in P122 steel in separate experimental

studies. Based on the experimental DOE study and these separate experimental studies

we could develop the formulae including comprehensive factors for predicting A1

temperature in P91 and P92 steels.

62
4.2 Sample Preparation

4.2.1 Materials

Test samples of 25g of various alloy compositions were made by tungsten arc

button melting in argon environment. The test samples were made by adding alloying

elements to particular commercial heats of P91, P92 and P122 steels, which were chosen

as base materials. Carbon was added to the test samples in form of 1080 steel wire. The

other elements were added in form of pure elements: manganese pieces, silicon granules,

chromium pieces, nickel wire, copper wire, molybdenum wire, tungsten wire, vanadium

wire and iron granules, their purities are shown in Table 4.4. The compositions of

specified P91, P92 and P122 steels base materials and 1080 steel wire are shown in Table

4.5.

Element Cr Mn Mo Ni Si V Cu W Fe
Purity 99% 99.9% 99.95 99.98% 99.999% 99.8% 99.9% 99.95% 99.98%
Table 4.4: Purity of pure elements.

C N Cr Mn Mo Ni Si V P S Cu W
P91 0.099 0.055 8.42 0.31 0.93 0.13 0.24 0.18 0.009 0.001 0.14 -
P92 0.11 0.045 8.64 0.48 0.38 0.13 0.32 0.174 0.012 0.006 0.19 1.62
P122 0.12 - 11.37 0.36 0.28 0.29 0.48 0.29 0.016 0.001 0.05 1.44
1080 0.81 - - 0.55 - - 0.23 - 0.011 0.007 - -
Table 4.5: Composition of base materials and 1080 steel wire (wt.%).
63
4.2.2 Sample Preparation

In order to simulate the manufacturing process of commercial P91, P92 and P122

steels, the procedures of sample melting and heat treatments were designed as follows:

1. Based on sample size of 25g, the weights of the base metal (P91 or P92 steel) and

of the pure elements needed to produce a sample with particular composition were

calculated.

2. The base metal and pure elements needed for each particular test sample were

weighted by an electronic balance (Mettler Toledo AB135-S), within 0.001g of

target weight.

3. The charge of each test sample (base metal and pure elements) was melted by a

tungsten arc button melting system in argon environment as shown in Figure

4.8(a). The charge was melted into a button sample over a copper hearth. The

melting arc current was 300A. To achieve homogeneous distribution of alloying

elements each button sample was flipped over and re-melted for two more times.

4. In the next step, the button sample was re-melted over an open copper hearth and

cast into a copper mold to produce a test sample with cubic shape, Figure 4.8(b).

The copper mold and the sample are shown in Figure 4.8(c) and Figure 4.8(d)

correspondingly. The casting was performed utilizing the same gas-tungsten arc

melting system in argon environment. The gas-tungsten arc current was 300A.

5. The test samples were homogenized at 1050˚C for 4 hours, and then cooled in the

furnace to room temperature.

64
6. The homogenized samples were hot-rolled at 1000˚C for four times and the

sample thickness was reduced 10% each time. After each hot-rolling step the

samples were reheated in the furnace to the rolling temperature. The total

thickness reduction was 34.4%. The sample shape after hot rolling is shown in

Figure 4.8(e).

7. The hot-rolled samples were normalized at 1050˚C for 0.5 hour. The samples

were heated with rate of 600˚C/h and then cooled in the air to room temperature.

The thermal history of the normalizing process is shown in Figure 4.8(f).

8. As shown in Figure 4.8(g), the normalized samples were tempered at 750˚C for 2

hours. Both the heating and the cooling rates were 200˚C/h.

65
a) b)

c) d) e)

f) f) g)

Figure 4.8: Schematics of sample melting, casting and heat treatment: a) Tungsten arc button
melting system; b) Sample casting system; c) Cubic copper casting mold; d) Shape of sample
after casting; e) Shape of sample after hot-rolling; f); Heat history of normalizing; g) Heat
history of tempering.

66
4.3 Evaluation of Composition Control

The chemical analyses of heat treated samples have been performed by NSL

Analytical. The chemical analysis of elements except for nitrogen was performed by

LECO Glow Discharge Spectrometer, and the concentration of nitrogen was measured by

LECO Furnace. As shown in Table 4.6, the chemical analysis results showed that carbon,

nitrogen and silicon were lost during the melting process, and even more during heat

treatment process. The loss of the metallic elements was negligible.

Energy Dispersive X-ray Spectroscopy (EDX) map analysis was performed by

SEM (Quanta 200). The magnification was 600, resolution was 128*128, and frame was

64. As shown in Figure 4.9, the EDX map analysis showed that all added metallic

elements were uniformly distributed inside the heat treated sample.

67
Cu Mo V

Cr W Si

Figure 4.9: EDX map analysis result of heat treated samples.

Element Sample 1 Sample 1 Sample 2 Sample 2 melted and


melted heat treated
C 0.104 0.064 0.104 0.034
Mn 0.7 0.64 0.711 0.716
Si 0.504 0.48 0.124 0.097
Cr 12.563 12.63 12.595 12.525
Ni 0.104 0.11 0.52 0.511
Cu 0.1 0.098 0.104 0.104
Mo 0.258 0.24 0.685 0.177
W 1.505 1.46 2.07 2.121
V 0.152 0.14 0.31 0.321
N 0.024 0.016 0.022 0.013
Table 4.6: Compositions of melted and heat treated samples (wt.%).

68
4.4 Determination of the A1 temperature

The critical temperatures were determined by the SS-DTA technique as shown in

Figure 2.28 [54]. The samples were heated at 600°C/h from 20°C to 650°C, and then at

28°C/h from 650 °C to 950°C (in accordance with ASTM A1033-04) in argon

atmosphere in a resistive heated convection furnace. Special Limit of Error type K

thermocouples (0.4% measurement error) were used for the temperature measurements.

The heating rate of the furnace is controlled by separate thermocouple of the furnace

control system, which is different from the thermocouple for sample measurement.

The SS-DTA technique was used to determine the thermal effect of reversion of

martensite to austenite and its starting and finishing temperatures. From 650 °C to 950°C,

the furnace heating rate was 28°C/h. The reference heating curve which is free of phase

transformation thermal effects should be a straight line with gradient of 28°C/h. The

digitally acquired heating history of sample was compared with the reference heating

curve free of phase transformation, and the temperature difference (δT) was plotted as a

function of temperature (T). There are two peaks in the plot of δT shown in Figure 4.10.

The first one corresponds to the ferro-magnetic to paramagnetic transformation (Curie

temperature A2). The second peak indicates the austenitization process; the start and the

end of the deviation peak indicate the beginning and finish of austenitization

(correspondingly the A1 and A3 temperatures).

69
A(bar)-2 28C-hr
950

900
A3

Temp (deg.C)
850

X: 0.2493
Y: 807.3

800
A1
X: -0.3252
Y: 754.3

750

A2
-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1
δT (deg.C)

Figure 4.10: A2, A1, A3 temperatures in laboratory melted heat of P91 steel determined by
SS-DTA.

70
4.5 Validation of formula predicted A1 temperature

In order to validate the formula predicted A1 temperature, two lab made P91 steel

specimens were chosen to be normalized at 1050°C for 30min and tempered at both

10°C below and 15°C above the formula predicted A1 temperature for 30min and then

air cooled to room temperature. Hardness testing and microscopy study were performed

on these tempered specimens. If the specimen was tempered above the A1 temperature,

there should be austenite formed during the tempering process, after air cooling the

austenite would transform to martensite. So the specimen tempered above the A1

temperature was supposed to have higher hardness, and both tempered martensite and

fresh martensite in the microstructure. The specimen tempered below the A1 temperature

was supposed to have full tempered martensite in the microstructure. The composition

of these two specimens is shown in Table 4.7.

Element Specimen1 Specimen2


C 0.07 0.085
Mn 0.278 0.29
Si 0.207 0.368
Cr 7.88 8.361
Ni 0.146 0.154
Cu 0.151 0.14
Mo 0.862 0.917
V 0.177 0.198
N 0.021 0.018
Table 4.7: Composition and SS-DTA measured A1 temperature of P91 steel specimens
(wt.%).

71
Chapter 5 RESULTS AND DISCUSSION

5.1 Model Based DOE

The elements with significant influence on JMat-Pro predicted A1 temperature in

P91 steel were screened out. Least square fitting model of A1 temperature in P91 steel as

a function of alloying compositions was developed using JMP8. As shown in Figure 5.1,

there were almost no deviations between the JMat-Pro predicted A1 temperatures and

model predicted A1 temperatures. The R-square value of 0.999 shows that the reliability

of this least square fitting model is 99.9%. R-square is calculated by diving regression

sum of squares with total sum of squares, it predicts how well model predict value relate

to actual value. As shown in Figure 5.2, the least square fitting model showed the

alloying elements with significant influence on JMat-Pro predicted A1 temperature in P91

steel are Ni, Mn, Si, N, Cr, Mo, C, V. Ni and Mn had the most significant effects. As

shown in Figure 5.3, all significant alloying elements have linear relation with JMat-Pro

predicted A1 temperature in P91 steel. The ferrite stabilizing elements Cr, Mo, Si, V

increased JMat-Pro predicted A1 temperature in P91 steel; and the austenite stabilizing

elements Ni, Mn, C, N decreased JMat-Pro predicted A1 temperature. The influence of

Mo and V on A1 temperature in P91 steel were very small, and the effect of Nb on A1

temperature was negligible. As shown in Figure 5.4, all of the influence curves of certain

element to the predicted A1 temperature were parallel to their corresponding original

72
curves, which meant that the changing of one element did not influence the effect of

another element on the predicted A1 temperature in P91 steel. It was conclude that there

were no interactions between the significant alloying elements, so interactions were not

taken into account for our further experimental study. What should be mentioned is that

the all the relations between the predicted A1 temperature and the composition was found

based on the composition range defined according to ASTM standard for P91 steel. These

relations are not supposed to be applicable to general steels with different composition

range.

Figure 5.1: The JMat-Pro predicted A1 temperature versus A1 temperature predicted by


the model based DOE.

73
Figure 5.2: Elements with significant influence on the JMat-Pro predicted A1 temperature
in P91 steel and their estimated effects.

Figure 5.3: Prediction profiler of model based DOE model of A1 temperature in P91 steel.

74
Figure 5.4: Interaction profiler of model based DOE model of A1 temperature in P91
steel.

75
5.2 Experimental Results

5.2.1 Measurement based DoE for P91 Steel

The results of the model based DoE study show that Ni, Mn, Si, N, Cr, Mo, C and

V have significant influence on A1 temperature in P91 steel. The concentration of C, N

and Si in the test samples could not be controlled as accurately as for the metallic

elements. For this reason only Ni, Mn, Cr, Mo, V and Cu were chosen as the factors of

measurement based DOE model. The composition range of experiment based DOE for

P91 steel is shown in Table 5.1. The original set point of Si, C and N for all samples were

selected correspondingly as 0.22 wt.%, 0.15 wt.% and 0.046 wt.%, assuming a loss to

almost same extent for these elements during sample preparation process. Chemical

analysis result in Table 5.2 shows Si, C and N became around 0.18 wt.%, 0.115 wt.% and

0.019 wt.% respectively after melting and heat treatment. It was found that commercial

P91 steels contain up to around 0.2 wt.% copper, although copper is not included in the

ASTM specification. The heat of P91 steel used as base material contained 0.14 wt.% Cu

and during sample preparation Cu was diluted farthest to 0.12 wt.%. It was assumed that

the effect of Cu on A1 temperature is negligible at copper concentration lower than 0.12

wt.% Cu. The measurement based DOE was developed using JMP8.

As shown in Figure 5.5, experimentally measured A1 temperature was defined as

response for measurement based DOE, designated as A1 in JMP8. Based on the analysis

above, according to model based DOE study results and experimental feasibility, the

factors for measurement based DOE were defined as shown in Figure 5.6. As shown in

Figure 5.7, in order to achieve a model with high accuracy both linear and quadratic

76
terms were defined for the experimental model. Based on JMat-Pro based DOE study

results, interaction terms were not taken into account. Two center points and one replicate

were defined, as shown in Figure 5.8. A measurement based DOE table was generated

Figure 5.9 using JMP8 that included a series of thirty six samples with different

compositions inside the ASTM specification range. Thirty six alloy samples were made

by button melting and then subjected to the homogenization, rolling and heat treatment

procedures described in Section 4.2.2. The A1 temperatures of these thirty six samples

measured using SS-DTA were entered in the response column for A1 in the Figure 5.9.

The testing error of type K thermocouple is ± 0.4%. However, specific data of the A1

temperatures was need for the JMP8 to develop the predictive formula, so the testing

error of type K thermocouple goes into the error of the predictive formula.

Figure 5.5: Defining the response for measurement based DOE in JMP8.

77
Figure 5.6: Defining the factors for measurement based DOE in JMP8.

Figure 5.7: Defining the model for measurement based DOE in JMP8.

78
Figure 5.8: Defining the number of center points and replicates for measurement based
DOE in JMP8.

Element Specification DOE


C 0.08-0.12 0.115
Cr 8.00-9.50 8.00-9.50
Cu - 0.117-0.40
Mn 0.30-0.60 0.304-0.60
Mo 0.85-1.05 0.775-1.05
N 0.03-0.07 0.019
Ni <0.40 0.108-0.40
Si 0.20-0.50 0.18
V 0.18-0.25 0.15-0.25
Table 5.1: Composition range of measurement based DoE for P91 steels (wt.%).

79
Figure 5.9: Measurement based DOE table with measured A1 temperature
values (testing error of type K thermocouple is ± 0.4%).

80
Based on the experimental DOE table with measured A1 temperature values,

statistical analysis was performed using JMP8. Least square fitting model of A1

temperature in P91 steel as a function of alloying compositions was developed using

JMP8, as shown in Figure 5.10. As shown in Figure 5.11, the R-square value of this

model is 0.843, which means the reliability of this least square fitting model is 84.3%. As

shown in Figure 5.12, compositions of Ni, Mn, Cr, Mo, V and Cu have significant effects

on A1 temperature in P91 steel. For Ni, Cu and V, influence of both linear term and

quadratic term were found to be significant. Linear terms of Mo and Cr, and quadratic

term of Mn were believed to be insignificant terms, because their P >│t│ value were too

high. Probabilities around or less than 0.05 are often considered as significant evidence

that the corresponding term is significant. As shown in Figure 5.13, Mn had almost linear

effect on A1 temperature in P91 steel, the other alloying elements had quadratic effects on

A1 temperature in P91 steel, specially the carbide formers Cr, Mo, V.

81
Figure 5.10: Predictive model of the A1 temperature in P91 steel as a function of
composition of Ni, Mn, Cr, Mo, V and Cu developed using JMP8 by
measurement based DOE study

82
Figure 5.11: The actual (measured) A1 temperature versus the experiment based DOE
model predicted A1 temperature

Figure 5.12: Elements with significant influence on measured A1 temperature in P91 steel
and their estimate effects.

83
Figure 5.13: Prediction profiler of measurement based DOE model of A1 temperature in
P91 steel.

A least square fitting model of the A1 temperature in P91 steel as a function of

compositions of metallic elements was developed, Equation (1).This model was based on

calculated compositions of test samples. Linear terms of Mo and Cr, and quadratic term

of Mn were deleted because they were believed to be insignificant terms.

A1 (˚C) = 814.2 – 3.61*((Mn-0.452)/0.148) + 9.38*((Cr-8.75)/0.75)2 – 7.79*((Ni-


0.254)/0.146)2 - 4.76*((Ni- 0.254)/0.146) - 3.89*((Cu-0.2585)/0.1415)2 –
2.34*((Cu-0.2585)/0.1415) + 3.10*((Mo-0.9125)/0.1375)2 – 1.01*((Mo-
0.9125)/0.1375) + 3.31*((V- 0.2)/0.05)2 -2.11*((V-0.2)/0.05) (1)

The results of chemical analysis performed by Materials Solution, Inc. (MSI) on

10 of the 65 measurement based DOE samples are compared to calculated composition of

these samples in Table 5.2. The average and standard deviation of deviations between

measured composition and calculated composition were calculated.

84
C Mn Si Cr Ni Cu Mo V N

Measured 0.12 0.29 0.18 7.97 0.43 0.12 0.98 0.23 0.021
2
Calculated 0.150 0.304 0.219 8.749 0.401 0.117 1.055 0.202 0.046

Measured 0.11 0.29 0.18 8.09 0.41 0.11 1 0.23 0.019


32
Calculated 0.149 0.304 0.219 8.769 0.403 0.117 1.055 0.202 0.046

Measured 0.11 0.45 0.18 8.24 0.24 0.27 0.87 0.24 0.020
6
Calculated 0.150 0.451 0.219 8.749 0.255 0.263 0.913 0.205 0.046

Measured 0.11 0.44 0.18 8.28 0.25 0.27 0.84 0.23 0.019
15
Calculated 0.150 0.454 0.219 8.740 0.259 0.263 0.912 0.201 0.046

Measured 0.12 0.6 0.18 8.09 0.24 0.27 0.99 0.23 0.019
10
Calculated 0.150 0.600 0.218 8.720 0.258 0.262 1.049 0.204 0.046

Measured 0.12 0.58 0.18 8.15 0.24 0.26 0.99 0.23 0.017
20
Calculated 0.150 0.597 0.219 8.747 0.255 0.263 1.048 0.197 0.046

Measured 0.12 0.59 0.16 7.31 0.4 0.4 0.86 0.23 0.019
18
Calculated 0.150 0.602 0.218 7.991 0.398 0.403 0.914 0.203 0.046

Measured 0.11 0.59 0.17 7.35 0.4 0.4 0.87 0.23 0.016
30
Calculated 0.150 0.603 0.219 7.997 0.399 0.399 0.912 0.201 0.046

Measured 0.12 0.44 0.18 9.09 0.24 0.41 1.02 0.29 0.020
22
Calculated 0.149 0.452 0.219 9.506 0.255 0.402 1.053 0.250 0.046

Measured 0.11 0.45 0.17 8.99 0.24 0.4 1.01 0.29 0.021
28
Calculated 0.151 0.455 0.219 9.497 0.257 0.400 1.052 0.251 0.046
Average of
Deviations -0.035 -0.010 -0.043 -0.591 -0.005 0.002 -0.053 0.031 -0.027
STD* of
Deviations 0.005 0.006 0.007 0.108 0.014 0.005 0.013 0.005 0.002
Table 5.2: Chemical analysis of measurement based DOE samples (wt.%).
*STD – Standard Deviation

85
Average and standard deviation values of concentration loss show that C, N and

Si were lost to almost same extent as we assumed. It was found that loss of all metallic

elements except for Cr was negligible. Cr was lost around 0.6 wt.%, standard deviation of

loss of Cr was 0.1 wt.%. Around 0.6 wt.% Cr was lost in all measurement based DOE

samples. The Cr term 9.38*((Cr-8.75)/0.75)2 of calculated composition based Equation (1)

was adjusted to 9.38*((Cr-8.75-0.6)/0.75)2, as shown in Equation (2). Equation (2) was

based on measured composition of test samples; the application range of this model is

shown in Table 5.3.

A1 (˚C) = 814.2 – 3.61*((Mn-0.452)/0.148) + 9.38*((Cr-8.15)/0.75)2 – 7.79*((Ni-


0.254)/0.146)2 - 4.76*((Ni- 0.254)/0.146) - 3.89*((Cu-0.2585)/0.1415)2 –
2.34*((Cu-0.2585)/0.1415) + 3.10*((Mo-0.9125)/0.1375)2 – 1.01*((Mo-
0.9125)/0.1375) + 3.31*((V- 0.2)/0.05)2 -2.11*((V-0.2)/0.05) (2)

Application
Element Specification Range
C 0.08-0.12 0.115
Cr 8.00-9.50 7.3-9.1
Cu - 0.12-0.40
Mn 0.30-0.60 0.30-0.60
Mo 0.85-1.05 0.8-1.0
N 0.03-0.07
Ni <0.40 0.10-0.40
Si 0.20-0.50 0.18
V 0.18-0.25 0.21-0.28
Table 5.3: Application range of Equation (2) (wt.%).

86
5.2.2 Determination of the effect of carbon and sillicon on the A1 temperature in P91

Steel

Effects of C and Si in P91 steel on A1 temperature were studied separately,

because C and Si were lost during the sample melting and heat treatment process, so the

concentration of C and Si could not be controlled as exact as metallic elements. The

composition of P91 steel used as base material is shown in Table 4.5.

C was added into samples in form of 1080 carbon wire, the composition of 1080

carbon wire is also shown in Table 4.5. Six samples with various C compositions

between 0.01~0.09 wt.% were made by button melting system. Chemical analysis of

these six samples was performed by MSI, as shown in Table 5.4. The standard deviation

values of compositions showed that the loss of other elements was negligible. The plot of

SS-DTA detected A1 temperatures to C concentration of corresponding samples (after

melting and heat treatment) is shown in Figure 5.14. The concentration of carbon

influences A1 temperature in a linear way. A linear fitting model was developed using

JMP8, as shown in Figure 5.14. The R-square value of this linear model is 0.926, which

means the reliability of this model is 92.6%, and the coefficient of C was -242.9.

Eight samples with various Si compositions between 0.20~0.50 wt.% were

produced by adding pure Si granules into P91 base material. The result of chemical

analysis by MSI is shown in Table 5.5. The standard deviation values of compositions

showed that the loss of other elements was negligible. As shown in Figure 5.14, a linear

relation was found between the concentration of Si and A1 temperature in P91 steel. A

87
linear model was developed using JMP8. The R-square value of the linear model is 0.930,

and the coefficient of Si is 87.5, as shown in Figure 5.14.

Based on separate experimental study of C and Si for P91 steel, two linear models

of A1 temperature in P91 steel as a function of concentration of C (3) and Si (4) were

develop using JMP8:

A1 (˚C) = 832.8 – 242.9*C (3)

A1 (˚C) = 803.5 + 87.5*Si (4)

Figure 5.14: Influence of C and Si on the A1 temperature in P91 steel.

88
Sample C1-1 C1-2 C2-1 C2-2 C3-1 C3-2 STD*
C 0.013 0.018 0.042 0.051 0.084 0.089 -
Cr 8.480 8.492 8.479 8.500 8.586 8.554 0.044
V 0.213 0.212 0.212 0.212 0.213 0.204 0.003
Mn 0.269 0.274 0.285 0.285 0.294 0.284 0.009
Mo 0.932 0.947 0.945 0.946 0.946 0.935 0.007
N 0.013 0.016 0.013 0.021 0.015 0.013 0.003
Ni 0.172 0.174 0.190 0.203 0.181 0.196 0.012
Si 0.244 0.251 0.252 0.244 0.272 0.286
0.017
Table 5.4: Measured composition of P91 steel samples with varied C concentration
(wt.%).
*STD – Standard Deviation

Sample P91-1 P92-2 Si3-1 Si3-2 Si2-1 Si2-2 Si1-1 Si1-2 STD*
C 0.077 0.078 0.089 0.076 0.084 0.085 0.081 0.080 0.004
Cr 8.294 8.333 8.295 8.294 8.310 8.361 8.277 8.271 0.030
Cu 0.146 0.133 0.153 0.157 0.145 0.140 0.148 0.146 0.007
Mn 0.280 0.274 0.265 0.279 0.275 0.290 0.270 0.269 0.008
Mo 0.909 0.914 0.914 0.904 0.892 0.917 0.902 0.901 0.008
N 0.020 0.017 0.016 0.016 0.018 0.018 0.016 0.016 0.001
Ni 0.144 0.157 0.162 0.151 0.150 0.154 0.149 0.159 0.006
Si 0.208 0.216 0.272 0.284 0.382 0.368 0.479 0.470 -
V 0.18 0.188 0.185 0.185 0.191 0.198 0.188 0.178 0.006
Table 5.5: Measured composition of P91 steel samples with varied Si concentration
(wt.%).
*STD – Standard Deviation

89
5.2.3 Determination of the effect of nitrogen on the A1 temperature in P91 Steel

Losing of N during melting and heat treatment process could not be controlled

and N could not be added into samples in our lab. The effect of N on A1 temperature in

P91 steel was determined by comparing the SS-DTA detected A1 temperature of melted

and heat treated P91 steel samples (P91-1 and P91-2) to the original commercial heat of

P91 steel. The chemical analysis by MSI (Table 5.6) showed that certain amount of C, N,

Si and small amount of metallic elements were lost during the melting and heat treatment

process. The SS-DTA detected A1 temperatures of these 3 samples are also shown in

Table 5.6. By combining models (2) (3) (4), the effects of the other elements on A1

temperature except for N could be calculated. The effect of N was determined as the

difference in A1 temperatures minus the effect of the other alloying elements. The A1

temperature was assumed to a have linear relation with the concentration of N. The

calculated coefficient for N is around -174.

Sample P91-1 P91-2 P91


C 0.077 0.078 0.099
Cr 8.294 8.333 8.42
Cu 0.146 0.133 0.14
Mn 0.28 0.274 0.31
Mo 0.909 0.914 0.93
N 0.02 0.017 0.055
Ni 0.144 0.157 0.13
Si 0.208 0.216 0.24
A1/°C 820.2 821.5 809
Table 5.6: Composition of commercial and laboratory heats of P91 steel samples used for
determination of the effect of nitrogen on the A1 temperature (wt.%).

90
5.2.3 Determination of the effect of molybdenum and tungsten on the A1 temperature

in P92 Steel

From the specification of P91 and P92 steels, the composition range of all

alloying elements in both steels is the same, except for Mo and W. The specification for

Mo is 0.30-0.60 wt. % in P92 steel and 0.85-1.05 wt. % in P91 steel, and there is no W in

P91 steel. The effects of Mo and W on A1 temperature in P92 steel were studied

separately. The composition of P92 steel used as base material is shown in Table 4.5, Mo

and W were added into P92 steel in form of pure wire.

Eight P92 steel samples with various W concentrations have been produced. The

chemical analysis of these samples was performed by MSI. The standard deviation values

of the composition showed that the loss of other elements was negligible, Table 5.7. The

A1 temperatures in these samples were determined by SS-DTA and plotted as a function

of concentration of W in Figure 5.15. A quadratic fitting model was developed using

JMP8, as shown in Figure 5.15. The R-square value of this model is 0.878, which means

the reliability of this model is 87.8%.

Six samples with various Mo compositions between 0.40-0.65 wt.% were

produced. The chemical analysis in these samples was performed by MSI, The standard

deviation values of the composition showed that the loss of other elements was negligible,

Table 5.8. The concentration of Mo had linear relation with A1 temperature in P92 steel,

as shown in Figure 5.15. A linear fitting model was developed using JMP8. The R-square

value of this linear model is 0.900, and the coefficient of Mo was 38.1.

91
Based on separate experimental study of W and Mo for P92 steel, a quadratic

model of A1 temperature in P92 steel as a function of concentration of W (5), and a linear

models of A1 temperature in P92 steel as a function of concentration of Mo (6) were

developed using JMP8:

A1 (˚C) = 860 – 6.34*W – 305.76*(W-1.8065)2 (5)

A1 (˚C) = 812 + 38.1*Mo (6)

Figure 5.15: Influence of W and Mo on the A1 temperature in P92 steel.

92
Element P92-1 P92-2 W1-1 W1-2 W2-1 W2-2 W3-1 W3-2 STD*
C 0.084 0.087 0.079 0.085 0.089 0.088 0.087 0.085 0.003
Cr 8.505 8.645 8.527 8.518 8.513 8.533 8.505 8.497 0.048

W 1.587 1.562 1.703 1.752 1.900 1.891 2.04 2.017 -


Mn 0.436 0.447 0.427 0.442 0.424 0.446 0.436 0.432 0.008
Mo 0.396 0.395 0.383 0.392 0.395 0.384 0.393 0.393 0.005
N 0.024 0.020 0.018 0.021 0.020 0.021 0.018 0.016 0.002
Ni 0.160 0.169 0.170 0.170 0.166 0.164 0.181 0.174 0.006
Si 0.306 0.310 0.303 0.306 0.298 0.299 0.314 0.297 0.006
V 0.151 0.154 0.154 0.16 0.19 0.157 0.17 0.16 0.013
Table 5.7: Measured composition of P92 steel samples with varied W concentration
(wt.%).
*STD – Standard Deviation

Sample P92-1 P92-2 Mo1-1 Mo1-2 Mo2-1 Mo2-2 STD*


C 0.084 0.087 0.093 0.083 0.088 0.091 0.004
Cr 8.505 8.645 8.555 8.481 8.547 8.487 0.061
V 0.151 0.154 0.158 0.151 0.159 0.165 0.005
Mn 0.436 0.447 0.438 0.423 0.427 0.425 0.009
Mo 0.396 0.395 0.506 0.507 0.628 0.634 -
N 0.024 0.020 0.022 0.019 0.021 0.020 0.002
Ni 0.160 0.169 0.161 0.170 0.162 0.161 0.004
Si 0.306 0.310 0.326 0.303 0.297 0.303 0.010
Table 5.8: Measured composition of P92 steel samples with varied Mo concentration.
*STD – Standard Deviation

93
5.2.4 Determination of the effect of copper on the A1 temperature in P122 Steel

As shown in Table 4.1, the specification for Cu in P122 steel is 0.30-1.70 wt.%.

P122 was the first Cr steel designed with high content of Cu. The influence of Cu on A1

temperature in P122 steel could be significant because of large composition range. Ten

P122 steel samples with various Cu compositions between 0.04~1.73 wt.% were made by

button melting system. The chemical analysis of these ten samples was performed by

MSI. As shown in Table 5.9, the standard deviation values of the composition show the

loss of alloying elements was negligible. The plot of SS-DTA detected A1 temperatures

to Cu concentration of corresponding samples is shown in Figure 5.16. At the range of

0.04~1.73 wt.%, Cu reduced the A1 temperature from around 835°C to 815°C. The

concentration of Cu influences the A1 temperature in a linear way. A linear fitting model

for the influence of Cu on the A1 temperature in P122 steel was developed using JMP8,

Figure 5.16. The R-square value of this linear model is 0.950, which means the reliability

of this model is 95.0%, and the coefficient of Cu was -12.4.

The specification for Cr in P122 steel is 10.00-12.50 wt.%, as shown in Table 4.1.

The heat of P122 steel used as base material contained 11.37 wt.% Cr. Since Cr in this

heat of P91 steel could only be diluted farthest to around 11 wt.%, test samples with

varying Cr compositions between 10 to 11 wt.% could not be made. A heat of P122 steel

with around 10 wt.% Cr needs to be collected in the future for determining the effect of

Cr on the A1 temperature in P122 steel.

94
Figure 5.16: Influence of Cu on the A1 temperature in P122 steel.

Sample C Mn Si Cr Ni Cu Mo W V

P122-1 0.10 0.33 0.49 11.50 0.31 0.04 0.28 1.65 0.25

P122-2 0.10 0.32 0.48 11.50 0.31 0.04 0.28 1.65 0.25

Cu1-1 0.10 0.33 0.48 11.40 0.31 0.29 0.28 1.66 0.24

Cu1-2 0.11 0.32 0.49 11.50 0.31 0.29 0.28 1.66 0.25

Cu2-1 0.10 0.32 0.49 11.40 0.31 0.81 0.28 1.65 0.25

Cu2-2 0.10 0.32 0.51 11.10 0.30 0.82 0.29 1.68 0.25

Cu3-1 0.10 0.32 0.47 11.30 0.31 1.28 0.27 1.65 0.25

Cu3-2 0.09 0.32 0.48 11.40 0.31 1.28 0.28 1.64 0.25

Cu4-1 0.10 0.32 0.48 11.30 0.31 1.75 0.27 1.62 0.24

Cu4-2 0.11 0.33 0.47 11.40 0.31 1.73 0.27 1.67 0.25

STD* 0.006 0.005 0.012 0.123 0.003 0.658 0.006 0.016 0.004
Table 5.9: Measured composition of P122 steel samples with varied Cu concentration.
*STD – Standard Deviation

95
5.3 Measurement based DOE formulae for the A1 Temperatures in CSEF Steels

5.3.1 Predictive formula for the A1 temperature in P91 Steel

Combining equations (2), (3), (4) and the coefficient of N derived in section 5.2.3,

a formula for predicting the A1 temperature in P91 steel that contain 0.12-0.40 wt.% Cu

was developed:

A1 (˚C) = 2113 - 243*C - 174*N + 87.5*Si - 24.4*Mn + 16.676*Cr2 – 271.81*Cr -


365.45*Ni2 + 153.05*Ni + 164*Mo2 - 299.2*Mo + 1320*V2 - 570*V - 194.3Cu2 +
83.9*Cu (7)

For P91 steel containing less than 0.12 wt.% Cu, the influence of Cu on A1

temperature in P91 steel was assumed to be negligible. The proposed formula for

predicting A1 temperature in P91 steel with less than 0.12% Cu is:

A1 (˚C) = 2120 - 243*C - 174*N + 87.5*Si - 24.4*Mn + 16.676*Cr2 – 271.81*Cr -


365.45*Ni2 + 153.05*Ni + 164*Mo2 - 299.2*Mo + 1320*V2 - 570*V (8)

A1 temperatures of ten commercial P91 steels determined by dilatometry

measurements were collected to validate the proposed formula of the A1 temperature in

P91 steel. The composition of these ten commercial P91 steels and corresponding

measured (by dilatometry) and formula predicted A1 temperature are shown in Table

5.10. The deviations of predicted from measured A1 temperatures are shown in Figure
96
5.17. The deviation range is between -3°C and 3°C. The concentration of Mn and Ni is

also shown in this diagram. The Mi+Ni content was believed to be a determining factor

for the A1 temperature in P91 and P92 steels. Steels containing higher combined amounts

of Ni and Mn are expected to have lower A1 temperature. The results of our study have

shown that the Mi+Ni content is not the only determining factor for A1 temperature in

P91 steel, and that the other alloying elements also have significant influence.

825 1
0.9
820
0.8
815 0.7

810 0.6
Dilatometry
0.5
805 Predictive Formula
0.4
Ni+Mn
800 0.3
0.2
795
0.1
790 0
1 2 3 4 5 6 7 8 9 10

Figure 5.17: Comparison of A1 temperature in P91 steels predicted by measurement


based DOE formula and A1 temperature determined by dilatometry.

97
C N Mn Si Cr Ni Cu Mo V Tested Predict Dev
A1/°C A1/°C
1 0.099 0.055 0.31 0.24 8.42 0.13 0.14 0.93 0.18 809 811 2

2 0.11 0.037 0.40 0.29 8.48 0.14 0.13 0.96 0.20 816 814 -2

3 0.095 0.045 0.39 0.30 8.53 0.18 0.03 0.92 0.18 821 820 0

4 0.11 0.054 0.41 0.31 8.59 0.16 0.15 0.99 0.20 813 816 3

5 0.11 0.058 0.40 0.28 8.47 0.18 0.06 0.99 0.19 814 812 -2

6 0.11 0.060 0.42 0.29 8.53 0.20 0.19 1.00 0.20 813 814 1

7 0.11 0.039 0.49 0.33 8.39 0.20 0.17 1.00 0.21 817 818 1

8 0.095 0.044 0.40 0.32 8.69 0.29 0.05 0.95 0.19 819 822 3

9 0.1 0.048 0.48 0.35 8.43 0.37 0.18 0.94 0.21 811 811 -3

10 0.11 0.051 0.51 0.30 8.43 0.38 0.15 0.92 0.22 801 801 0

Max 0.11 0.060 0.52 0.35 8.69 0.38 0.19 1.00 0.22 821 822 -3

Min 0.095 0.037 0.31 0.24 8.39 0.13 0.03 0.92 0.18 801 801 3

Table 5.10: Composition (wt.%) of 10 commercial P91 steels and corresponding


measured (by dilatometry) and formula predicted A1 temperature.

The predicted A1 temperatures for the ten P91 steel samples by developed

predictive formula, by JMatPro and by the expression developed by Santella [65] are

shown in Table 5.11. The data are plotted in Figure 5.18; our predictive formula has the

closest agreement with measured values by dilatometry. The reasons for the significant

deviation between the compared predictive methods require further clarification.

Validation by comparison to a larger database of experimentally determined values of the

A1 temperature in P91 steel is also suggested.

98
Predictive Santella's Jmat-
Sample Dilatometry Formula Expression Pro
1 809 811 829.2 808.5
2 816 814 821.5 807.5
3 821 820 820.3 806.7
4 813 816 823.1 804.1
5 814 812 821.5 804.0
6 813 814 820 804.0
7 817 818 814 798.7
8 819 822 813.3 800.1
9 811 811 802 785.8
10 801 801 798.1 782.4
Table 5.11: Predictions for P91 steel samples from different available models (°C).

840 Unit: °C

830

820
Predictive Formula
810 Santella's Expression
Jmat-Pro
800
equal line
790

780
800 805 810 815 820 825
(Dilatometry Measurement/°C)

Figure 5.18: Predictions for P91 steel samples from different available models.

99
5.3.2 Predictive formula for the A1 temperature in P92 Steel

Combining equations (2), (5), (6) and the coefficient of N derived in section 5.2.3,

a formula for predicting A1 temperature in P92 steel that contain 0.12-0.40 wt.% Cu was

developed:

A1 (˚C) = 970.5 - 243*C - 174*N + 87.5*Si - 24.4*Mn + 16.676*Cr2 – 271.81*Cr -


365.45*Ni2 + 153.05*Ni + 38*Mo + 1320*V2 - 570*V - 194.3*Cu2 + 83.9*Cu -
305.7475*W2 + 1098.36*W (9)

For P92 steel containing less than 0.12 wt.% Cu, the influence of Cu on A1

temperature in P92 steel was assumed to be negligible. The proposed formula for

predicting A1 temperature in P92 steel with less than 0.12% Cu is:

A1 (˚C) = 976.5 - 243*C - 174*N + 87.5*Si - 24.4*Mn + 16.676*Cr2 – 271.81*Cr


- 365.45*Ni2 + 153.05*Ni + 38*Mo + 1320*V2 - 570*V - 305.7475*W2 +
1098.36*W (10)

The A1 temperatures measured by dilatometry in nineteen commercial P92 steels

are summarized in Figure 5.19. The composition of these nineteen commercial P92 steels

and corresponding measured (by dilatometry) and formula predicted A1 temperature are

shown in Table 5.12. The deviations range of predicted from measured A1 temperature is

100
between -11°C and 11°C. As in steel P91, the Mi+Ni content was not the only

determining factor for A1 temperature in P92 steel.

820 1
0.9
815
0.8
810 0.7

805 0.6
Dilatometry
0.5
800 Predictive Formula
0.4
Ni+Mn
795 0.3
0.2
790
0.1
785 0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19

Figure 5.19: Comparison of A1 temperature in P92 steels predicted by the proposed


measurement based DOE formula and A1 temperature determined by dilatometry.

The predictions for the A1 temperature in these nineteen P92 steel samples by

developed predictive formula, by JMatPro and by the expression developed by Santella

[65] are compared in Table 5.13 and in Figure 5.20. The predictive formula developed in

this work has the closest agreement with measured values by dilatometry.

The larger deviation range between the A1 temperature predicted by the formula

developed in this work and the dilatometry measurements, and among the compared

101
predictive methods requires further clarification. Validation towards a larger database of

experimentally determined A1 temperatures is necessary, since the currently available

database for P92 steel covers only a narrow range of Mn + Ni contents (Figure 5.19).

102
Test Formula
C Mn Si Cr Ni Cu Mo W V N A1/°C A1/°C Dev
0.11 0.48 0.32 8.64 0.13 0.19 0.38 1.62 0.17 0.045 812 803 -9
1
0.11 0.46 0.32 8.78 0.36 0.19 0.41 1.91 0.20 0.045 809 805 -4
2
0.1 0.45 0.34 8.72 0.37 0.16 0.4 1.69 0.19 0.042 807 808 1
3
0.11 0.47 0.37 8.87 0.37 0.19 0.41 1.72 0.21 0.045 813 812 -1
4
0.13 0.47 0.4 8.62 0.38 0.11 0.4 1.85 0.20 0.045 810 802 -8
5
0.1 0.47 0.34 8.68 0.38 0.15 0.39 1.92 0.20 0.043 814 803 -11
6
0.1 0.48 0.36 8.78 0.37 0.14 0.42 1.73 0.22 0.046 811 811 0
7
0.1 0.47 0.34 8.75 0.38 0.16 0.4 1.90 0.21 0.045 807 806 -1
8
0.1 0.49 0.29 8.68 0.37 0.19 0.39 1.73 0.19 0.040 798 804 6
9
0.09 0.49 0.35 8.81 0.37 0.11 0.41 1.81 0.19 0.037 804 815 11
10
0.11 0.49 0.33 8.77 0.37 0.14 0.4 1.78 0.19 0.045 813 807 -6
11
0.11 0.49 0.31 8.81 0.37 0.21 0.41 1.84 0.21 0.052 812 805 -7
12
0.12 0.49 0.28 8.66 0.37 0.17 0.39 1.77 0.21 0.047 805 797 -8
13
0.11 0.48 0.36 8.86 0.38 0.13 0.4 1.59 0.20 0.049 807 796 -11
14
0.13 0.49 0.35 8.99 0.38 0.14 0.39 1.86 0.19 0.048 811 806 -5
15
0.1 0.49 0.32 8.61 0.38 0.18 0.4 1.79 0.21 0.043 801 805 4
16
0.1 0.48 0.32 8.75 0.39 0.16 0.41 1.74 0.19 0.043 798 807 9
17
0.1 0.49 0.35 8.63 0.39 0.2 0.4 1.64 0.19 0.044 803 800 -3
18
0.11 0.49 0.33 8.66 0.39 0.18 0.4 1.81 0.21 0.046 798 802 4
19
0.13 0.5 0.4 8.99 0.39 0.2 0.42 1.92 0.22 0.049 814 815 11
Max
0.09 0.45 0.28 8.61 0.13 0.11 0.38 1.59 0.17 0.037 798 796 -11
Min
Table 5.12: Composition (wt.%) of 19 commercial P92 steels and corresponding
measured (by dilatometry) and formula predicted A1 temperature.

103
Predictive Santella's Jmat-
Dilatometry
Sample Formula Expression Pro
1 812 803 810 804.8
2 809 805 801 793.6
3 807 808 800 793
4 813 812 800 792.1
5 810 802 798 793.3
6 814 803 801 793.4
7 811 811 800 793.8
8 807 806 801 793.4
9 798 804 797 789
10 804 815 801 795.4
11 813 807 797 791
12 812 805 798 789.1
13 805 797 795 788.4
14 807 796 796 790.1
15 811 806 797 790.2
16 801 805 798 790.1
17 798 807 797 789.9
18 803 800 795 787.7
19 798 802 797 788.8
Table 5.13: Predictions for P92 steel samples from different available models (°C).

820 Unit: °C

815

810

805 Predictive Formula


Santella's Expression
800
Jmat-Pro
795 equal line

790

785
795 800 805 810 815 (Dilatometry Measurement/°C)

Figure 5.20: Predictions for P92 steel samples from different available models.

104
5.4 Validation of formula predicted A1 temperature

For validation of the A1 temperature predictions by the measurement-based DOE

formula, two P91 steel samples (specimen 1 and specimen 2) were collected. As shown in

Table 5.12, the formula predicted A1 temperatures of specimen 1 and specimen 2 are

826.5°C and 834.8°C correspondingly. Specimen 1 and specimen 2 were both tempered

at 825°C and 850°C for 30 min and then air cooled to room temperature. The macro-

hardness of tempered samples is shown in Table 5.12. As shown in Figure 5.19 (a) and

(d), the microstructure of specimens tempered at 825°C was tempered martensite, which

indicated they were not over tempered. As shown in Figure 5.19 (b) and (e), the

microstructure in samples tempered at 850°C contained both ferrite and fresh martensite.

This is a proof that the A1 temperature was exceeded and transformation to austenite

occurred during tempering. Because 850°C is between the A1 temperature and A3

temperature, the equilibrium phases at 850°C are ferrite and austenite. During the air

cooling fresh martensite formed from austenite. A1 temperature in specimen 1 is lower

than in specimen 2, so at 850°C more austenite formed in specimen 1. After air cooling,

more martensite formed in specimen 1. Figure 5.19 (b) shows more martensite in the

microstructure than Figure 5.19 (e), and the hardness of specimen 1 after tempering at

850°C was higher than of specimen 2. Figure 5.19 (c) and (f) show the micro-hardness of

the two phases in the microstructure: the softer white phase is ferrite and the harder lath

structure phase is martensite. The load of micro-hardness testing was 50g.

105
Specimen1 Specimen2
C 0.07 0.085
Mn 0.278 0.29
Si 0.207 0.368
Cr 7.88 8.361
Ni 0.146 0.154
Cu 0.151 0.14
Mo 0.862 0.917
V 0.177 0.198
N 0.021 0.018
Formula Predicted
826.5 834.8
A1/°C
Macro-hardness after
tempering at 825°C/HV 140 152
(testing load 1Kg)
Macro-hardness after
tempering at 850°C/HV 297 238
(testing load 1Kg)
Table 5.14: Composition and hardness of tempered P91 specimens.

106
a d

Cr1-1
Spe.1 Spe.2
825°C 825°C
825°C
140HV 152HV

b e

Spe.1 Spe.2
850°C 850°C
297HV 238HV

c f

323HV

364HV

193HV
173HV
Cr1-1
Spe.1 Spe.2
850°C 850°C
850°C
297HV 238HV

Figure 5.21: Microscopy of tempered samples (X1000): a) Specimen1 tempered at 825°C for
30min, hardness is 140 HV; b) Specimen1 tempered at 850°C for 30min, hardness is 297 HV;
c) Micro-hardness of specimen1 tempered at 850°C (load 50g); d) Specimen2 tempered at
825°C for 30min, hardness is 152 HV; e) Specimen2 tempered at 850°C for 30min, hardness
is 238 HV; f) Micro-hardness of specimen2 tempered at 850°C (load 50g).

107
Chapter 6 CONCLUSIONS

1. The influence of the alloying elements on the A1 temperature in creep strength

enhanced ferritic steels was investigated by model-based DOE and by

measurement based DOE. It was found that the elements with significant effect on

the A1 temperature in P91 steel are Ni, Mn, Si, N, Cr, Mo, C, V and Cu.

2. Nickel and manganese have a strong effect on the A1 temperature in P91 and P92

steels, but they are not the only determining factor. All the other alloying

elements influence the A1 temperature especially C, N, Si, Cr, and W.

3. The relations of the A1 temperature to the concentrations of strong carbide

formers (Cr, Mo, V in P91 and P92 steels and W in P92 steel) are quadratic

functions. The mechanism behind the quadratic effect of strong carbide formers

on the A1 temperature in P91 and P92 steels has not been clarified and needs

further investigation.

108
4. Predictive formulae for A1 temperature in P91 and P92 steels have been

developed using measurement based DOE. The proposed predictive formulae are

shown below:

For P91 steel containing 0.12-0.40 wt.% Cu:

A1 (˚C) = 2113 - 243*C - 174*N + 87.5*Si - 24.4*Mn + 16.676*Cr2 – 271.81*Cr

- 365.45*Ni2 + 153.05*Ni + 164*Mo2 - 299.2*Mo + 1320*V2 - 570*V - 194.3Cu2

+ 83.9*Cu

For P91 steel containing less than 0.12 wt.% Cu:

A1 (˚C) = 2120 - 243*C - 174*N + 87.5*Si - 24.4*Mn + 16.676*Cr2 – 271.81*Cr

- 365.45*Ni2 + 153.05*Ni + 164*Mo2 - 299.2*Mo + 1320*V2 - 570*V

For P92 steel containing 0.12-0.40 wt.% Cu:

A1 (˚C) = 970.5 - 243*C - 174*N + 87.5*Si - 24.4*Mn + 16.676*Cr2 – 271.81*Cr

- 365.45*Ni2 + 153.05*Ni + 38*Mo + 1320*V2 - 570*V - 194.3*Cu2 + 83.9*Cu

- 305.7475*W2 + 1098.36*W

For P92 steel containing less than 0.12 wt.% Cu:

A1 (˚C) = 976.5 - 243*C - 174*N + 87.5*Si - 24.4*Mn + 16.676*Cr2 – 271.81*Cr

- 365.45*Ni2 + 153.05*Ni + 38*Mo + 1320*V2 - 570*V - 305.7475*W2

+ 1098.36*W

109
5. The proposed predictive formulae have been validated by experimental

measurements. The deviation ranges of the predicted A1 temperatures from

dilatometry measured values in P91 and 19 P92 steels are correspondingly ±3°C

and ±11°C. This deviation range is slightly wider than the accuracy of

temperature measurement range that can be achieved by special limits of error

type K thermocouple (0.4% or +/-3.2°C at 800°C).

6. Further validation of the proposed predictive formulae is suggested by

comparison to a larger database of experimentally determined A1 temperatures

that covers a wider range of steel compositions. Following additional validation,

the proposed predictive formulae for could be used for development of proper

post welding heat treatment procedures for P91 and P92 steel welds.

7. Additional study of the effect of Cr on the A1 temperature in P122 steel is

necessary in order to develop a predictive equation for the A1 temperature in this

steel. It was found that copper has a strong effect on the A1 temperature in P122

steel. Varying the copper content in the composition specification range of P122

steel results in 20°C variation of the A1 temperature.

110
References

1. K.W. Andrew, “Empirical Formulae for the Calculation of Some Transformation


Temperatures”, Journal of The Iron and Steel Institute, Vol. 203, July, 1965, pp.
721-727.

2. S. Spigarelli and E. Quandrini, Analysis of the Creep Behavior of Modified P91


(9Cr-1Mo-NbV), Materials and Design, Vol. 23, 2002, pp. 547-552.

3. J. Hald, “Microstructure and Long-term Creep Properties of 9-12% Cr Steels”,


International Journal of Pressure Vessels and Piping, Vol. 85, 2008, pp. 30-37.

4. Michael Santella, “Advanced Pressure Boundary Materials”, The 22nd Annual


Conference on Fossil Energy Materials, National Energy Technology Laboratory,
8-10 July, 2008.

5. A. Iseda, Y. Sawaragi, K. Kato and F. Masuyama, “Development of a New 0.1C-


11Cr-2W-0.4Mo-1Cu Steel for Large Diameter and Thick Wall Pipe for Boilers”,
Proceedings of the Fifth International Conference on Creep of Materials, Lake
Buena Vista, Florida, 18-21 May, 1992, pp. 389-397.

6. Y. Tsuchiba and R. Yamaba, K. Tokuno, K. Hashimoto, T. Ogawa and T. Takeda,


“BOP Manufacturing and Properties of ASTM A 387 Grade 91 Steel Plates”,
Proceedings of The 1990 Pressure Vessels and Piping Conference, Nashville,
Tennessee, 17-21 June, 1990, pp. 105-114.

7. R. B. Scarlin, “Improved Materials for High Efficiency Steam Turbines”,


Advances in Turbine Materials, Design and Manufacturing: Proceedings of the
Fourth International Charles Parsons Turbine Conference, November, 1997, pp.
242-256.

8. H. Cerjak and E. Letofsky, “The Effect of Welding on The Properties of


Advanced 9-12%Cr Steels”, Science and Technology of Welding and Joining,
Vol. 1(1), 1996, pp. 36-42.

9. F. Masuyama, T. Yokoyama, Y. Sawaragi, and A. Iseada, “Development of


Tungsten Strengthened Low Alloy Steel with Improved Weldability”, Materials
for Advanced Power Engineering 1994: Proceedings of a Conference Held in
Liege, Belgium, October, 1994, pp. 173-181.

111
10. C. D. Laudin, P. Liu and Y. Cui, “A Literature Review on Characteristics of High
Temperature Ferritic Cr-Mo Steels and Weldments”, WRC bulletin 454, Welding
Research Council, August, 2000, pp. 1-36.

11. G. Guntz, M. Julien, G. Kottmann, F. Pellicani, A. Pouilly, and J. C. Vaillant,


“The T91 Book: Ferritic Tubes and Pipe for High Temperature Use in Boilers”,
Vallourec Industries, France, 1990.

12. Hisashi Naoi, Hiroyuki Mimura, Masahiro Ohgami, Mizuo Sakakibara, Satoshi
Araki, Yasuo Sogoh, Tadao Ogawa, Hideo Sakurai and Toshio Fujita,
“Development of Tubes and Pipes for Ultra-Supercritical Thermal Power Plant
Boilers”, Nippon Steel Technical Report, No.57, April, 1993, pp. 22-27.

13. Hiroyuki Mimura and Tetsuo Ishitsuka, “Boiler Pipes and Tubes for High
Efficiency of Power Generation”, Nippon Steel Technical Report, No.81, January,
2000, pp. 74-78.

14. Kent K. Coleman and W. F. Newell Jr., “P91 and Beyond: Welding the New-
generation Cr-Mo Alloys for High-Temperature Service”, Welding Journal,
August, 2007, pp. 29-33.

15. J. A. Francis, W. Mazur and H. K. D. H. Bhadeshia, “Type IV Cracking in


Ferritic Power Plant Steels”, Materials Science and Technology, Vol. 12(12),
2006, pp. 1387-1395.

16. A. Iseda, Y. Sawaragi, H. Teranishi, M. Kubota and Y. Hayase, “Development of


New 12%Cr Steel Tubing (HCM12) for Boiler Application”, The Sumitomo
Search, No. 40, November, 1989, pp. 41-56.

17. J. F. Copeland and G. J. Licina, “A Review of 2 1/4Cr-1Mo Steel for LMFBR


Steam Generator Application”, Symposium on Structural Materials for Service at
Elevated Temperatures in Nuclear Power Generation, ASME, Huston, 1975.

18. Fujio Ade, “Precipitate Design for Creep Strengthening of 9% Cr Tempered


Martensitic Steel for Ultra-supercritical Power Plants”, Science and Technology
of Advanced Materials, 9 (2008) 013002, March 13th, 2008.

19. M. Y. Wey, T. Sakuma and T. Nishizawa, “Growth of Alloy Carbide Particles in


Austenite”, Transactions of Japan Institute of Metals, Vol. 22(10), 1981, pp. 733-
742.

20. F. Abe, M. Taneike1, and K. Sawada, “Alloy Design of Creep Resistant 9Cr
Steel Using a Dispersion of Nano-sized Carbonitrides”, International Journal of
Pressure Vessels and Piping, Vol. 84(1-2), January-February, 2007, pp. 3-12.
112
21. John C. Lippold and Damian J. Kotecki, “Welding Metallurgy and Weldability of
Stainless Steels”, Wiley-Interscience, 2005, pp. 20.

22. Fujimitsu Masuyama, “History of Power Plant and Progress in Heat Resistant
Steels”, The Iron and Steel Institute of Japan(ISIJ) International, Vol. 41(6), 2001,
pp. 612-625.

23. N. Takahashi, T. Fujita and T. Yamada, Tetsu-to-Hagane, Vol. 61, 1975, pp.
2263-2273.

24. J. Hald, “Metallography and Alloy Design in The COST 536 Action”, Advanced
Materials for Power Engineering 2006: Proceedings of 8th Liege Conference,
Liege, Belgium, 18-20 September, 2006, pp. 917-930.

25. P. Peel, B. Scarlin and R. Vanstone, “From Materials Development to Advanced


Steam Turbines”, Proceedings of 7th International Charles Parsons Turbine
Conference, Glasgow, UK, 11-13 September, 2007, pp. 389.

26. K. Sawada, M. Taneike, K. Kimura and F. Abe, “Effect of Nitrogen Content on


Microstructural Aspects and Creep Behavior in Extremely Low Carbon 9Cr Heat-
resistant Steel”, The Iron and Steel Institute of Japan(ISIJ) International, Vol.
44(7), 2004, pp. 1243-1249.

27. Mats Hattetrand and Hans-Olof Andren, “Microstructure Development During


Aging of an 11% Chromium Steel Alloyed with Copper”, Materials Science and
Engineering: A, Vol. 318, 2001, pp. 94-101.

28. M. Ohgami, H. Minura, and H. Naoi, “Creep Rupture Properties and


Microstructures of a New Ferritic W Containing Steel”, Proceedings of the Fifth
International Conference on Creep of Materials, Lake Buena Vista, Florida, 18-21
May, 1992, pp. 69-73.

29. G. R. Prescott and C. F. Braun, “Update on Modified 2 1/4Cr-1Mo and 3Cr-1Mo


Alloys”, MPC, HPV-7, 1983.

30. K. Yoshikawa, A. Iseda, M. Yano, F.Masuyama, T. Daikoku and H. Haneda, First


International Conference on Improved Coal-Fired Power Plants, Palo Alto,
California, 1986.

31. Y. Tsuda, R.Ishii, M. Yamada, T. Azuma, Y. Tanaka and Y. Ikeda, Proceedings


of the International Conference on Power Engineering-’97, Japanese Society of
Mechanical Engineering, Tokyo, Vol. 2, 1997, pp. 131.

32. T. Fujita, Thermal and Nuclear Power, Vol. 42, 1991, pp. 1485.
113
33. Nippon Steel Corporation, “Data Package for NF616 Ferritic Steel (9Cr-0.5Mo-
1.8W-Nb-V)”, the second Edition, March, 1994.

34. T. Onizawa, T. Wakai, M. Ando and K. Anto, “Effects of Vanadium and Niobium
on Precipitation Behavior and Mechanical Properties of High Cr Steel”, Nuclear
Engineering and Design, Vol. 238, 2008, pp. 408-416.

35. K. Hidak, Y. Fukui, S. Nakamura, “Development of Heat Resistant 12%CrWCoB


steel Rotor for USC Power Plant”, Advanced Heat Resistant Steels for Power
Generation, Miramar Palace, San Sebastian, Spain, 27-29 April, 1998.

36. M. Miyazaki, M. Yamada, Y. Tsuda and R. Ishii, “Advanced Heat Resistant


Steels for Steel Turbines”, Advanced Heat Resistant Steels for Power Generation,
Miramar Palace, San Sebastian, Spain, 27-29 April, 1998.

37. D. Dulieu, K. W. Tupholme and G. J. Butterworth, “Development of Low-


activation Martensitic Stainless Steels”, Journal of Nuclear Materials, Vol. 143,
1986, pp. 1097-1101.

38. D. S. Gelles, “Effects of Irradiation on Low Activation Ferritic Alloys”, Reduced


Activation Materials for Fusion Reactors, ASTM STP 1047, Philadelphia:
American Society for Testing and Materials, 1990, pp. 113-129.

39. R. L. Klueh, D, J. Alexander and M. Rieth, “Effect of Tantalum on the


Mechanical Properties of a (Cr-2W-0.25V-0.07Ta-0.1C Steel”, Journal of Nuclear
Materials, Vol. 273, 1999, pp. 146-154.

40. H. Kjartansson and J. Hald, “Tantalum-containing Z-phase in 12%Cr Martensitc


Steels”, Scripta Materialia, Vol. 60, 2009, pp. 811-813.

41. Y. Wang, K. H. Mayer, A. Scholz, C. Berger, H. Chilukuru, K. Durst and W.


Blum, “Development of new 11%Cr Heat Resistant Ferritic Steels with Enhanced
Creep Resistance for Steam Power Plants with Operating Steam Temperature up
to 650°C”, Materials Science and Engineering A, Vol. 511-512, 2009, pp. 180-
184.

42. K. H. S Mayer, Proceedings of the 29th MPA Seminar Materials & Component
Behavior in Energy & Plant Technology Stuttgart, 9-10 October 2003, MPA
Stuttgart, Germany, 2003.

43. J. Piron Abellan, P. J.Ennis, L. Singheiser, W. J. Quadakkers, Material for


Advanced Power Engineering 2006: Proceedings of the 8th Liege Conference,
Forschungszentrum Julich, Germany, Vol. 3, 2006, pp. 1285-1295.

114
44. Y. Murata, M. Morinaga and R. Hashizume, “Development of Ferritic Steels for
Steam Turbine Rotors with the Aid of a Molecular Orbital Method”, Advances in
Turbine Materials, Design and Manufacturing: Proceedings of the Fourth
International Charles Parsons Turbine Conference, November, 1997, pp. 270-282.

45. K. Kimura, K. Sawada, H. Kushima and Y. Toda, “Influence of Delta Ferrite


Phase on Long-Term Creep Strength of High Chromium Ferritic Creep Resistant
Steels”, Proceedings of 7th International Charles Parsons Turbine Conference,
Glasgow, UK, 11-13 September, 2007.

46. E. Schnabel, P. Schwaab and H. Weber, Stahl u. Eisen, Vol. 107(2), 1987, pp. 691

47. John Hald and Hilmar K. Danielsen, “Z-phase Strengthened Martensitic 9-12%Cr
Steels”, Proceedings of 3rd Symposium on Heat Resistant Steels and Alloys
for High Efficiency USC Power Plants, National Institute for Materials Science,
Tsukuba, Japan, 2-6 June, 2009.

48. H. K. D. H. Bhadeshia, “Thermal Analysis


Techniques”, http://www.msm.cam.ac.uk/phasetrans/2002/Thermal1.pdf,
University of Cambridge, Material Science & Metallurgy, 2002.

49. M. Gojic, M. Suceska and M. Rajic, “Thermal Analysis of Low Alloy Cr-Mo
Steel”, Journal of Thermal Analysis and Calorimetry, Vol. 75, 2004, pp. 947-956

50. G. W. H. Hohne, W. F. Hemminger and H. J. Flammersheim, “Differential


Scanning Calorimetry”, 2nd Edition, Springer, 2003, pp. 10.

51. H. K. D. H. Bhadeshia, “Bainite: Overall Transformation Kinetics”, Journal De


Physique, Vol. 43, December, 1982, pp. C4-443.

52. M. Takahashi and H. K. D. H. Bhadeshia, “The Interpretation of Dilatometric


Data for Phase Transformation in Steels”, Journal of Materials Science Letters,
Vol. 8, 1989, pp. 477-478.

53. B.T. Alexandrov and J.C. Lippold, “In-Situ Weld Metal Continuous Cooling
Transformation Diagrams”, IIW Doc. IX-2162-05, pp. 1-16.

54. B. T. Alexandrov, J.K. Tatman, G. Murray and J.C. Lippold, “Non-equilibrium


Phase Transformation Diagrams in Engineering Alloys”, Trends in Welding
Research, Proceedings of the 8th International Conference, ASM International,
2009, pp. 467-476.

115
55. Overview of Instron ETMT, Center of Accelerated Maturation of Materials,
Materials Science and Engineering Department, Ohio State
University, http://www.camm.ohio-state.edu/Facilities/InstronETMT.html

56. P. Payson and C.H. Savage, “Martensite Reactions in Alloy Steels”, Transactions
of ASM, Vol. 33, 1944, pp. 261-275.

57. Discussion part of reference 56.

58. E.S. Rowland and S.R. Lyle, “The Application of Ms Points to Case Depth
Measurement”, Transactions of ASM, Vol. 37, 1946, pp.27-47.

59. R. A. Grange and H.M. Stewart, “The Temperature Range of Martensitic


Formation”, Transactions of AIME, Vol. 167, 1946, pp. 467-490.

60. W. Steven and A.G. Haynes, “The Temperature of Formation of Martensite and
Bainite in Low-alloy Steels”, Journal of The Iron and Steel Institute, Vol. 183,
1956, pp. 349-359.

61. C.Y. Kung and J.J. Rayment, “An Examination of the Validity of Existing
Empirical Formulae for the Calculation of Ms Temperatures”, Metallurgical
Transactions A, Vol. 13A, 1982, pp. 328-331.

62. Jicheng Zhao, “Continuous Cooling Transformations in Steels”, Material Science


and Technology, Vol. 8, November, 1992, pp. 997-1003.

63. L. Beres, Z. Beres and W. Irmer, Welding and Cutting, Vol. 8, 1994, pp. 128.

64. Zhenbo Zhao, Xin Guan, Chujie Wan, Cheng Liu and Derek O. Northwood, “A
Re-examination of the B0 and Bs temperatures of steel”, Materials and Design,
Vol. 21, 2000, pp. 207-209.

65. Michael Santella, “Influence of Chemical Compositions on Lower Ferrite-


Austenite Transformation Temperatures in 9Cr Steels”, Proceedings of the ASME
2010 Pressure Vessels and Piping Division/K-PVP Conference, Washington,
USA, July18-22, 2010.

116

Potrebbero piacerti anche